Anda di halaman 1dari 100

Why Take Chemistry, Anyway?

Head back to the helpdesk Find more tutorials Try some practice worksheets

One question I get a lot from students is "Why do we have to take chemistry? What's it good for in the real world?" Actually, that's two questions, but I'll answer them anyway. Chemistry is used in everything we do in modern society. The aspirin (or acetominophen, or ibuprofen, or naproxen sodium, or ketoprofen, etc.) you take when you have a headache was manufactured by chemists working for pharmaceutical companies. The gasoline you use to operate your car was formulated by petroleum chemists to give it the best possible operating properties. If you have contact lenses, the plastics used to make the lens, as well as the solutions you use to clean them were developed by chemists. No matter where you go, there's chemistry! But the question remains: If you're not going to be a chemist, why do you need to know how this stuff works? Shouldn't you just be happy that it works at all? Honestly, you don't really need to know the exact details of all the chemistry in your lives. Chances are that it would take you years to learn all the chemistry, and even then it wouldn't be all that handy to know. However, that's not my (or any other teacher's) goal in teaching you chemistry. I'm not trying to teach you everything - I'm just trying to give you enough of a background that you can understand some of the basics of what's going on around you. There are too many people trying to make money off the ignorance of people who haven't bothered to learn chemistry, and I don't want you to become one of them. An example: There are pills you can buy which advertise that they contain "life-giving oxygen." This sounds wonderful! Let's go buy some. However, you need to realize at the same time that water also contains "life-giving oxygen." If you're trying to get healthy, I'll only charge you half as much for a glass of water as these other guys are charging you for the pills. If you don't know any better, you might get roped in by that scam. I want you to learn chemistry because it helps you to realize what's good and what's garbage out in society. P.T. Barnum said that "There's a sucker born any minute." I don't want you to be that sucker.

The Scientific Method

Head back to the helpdesk Find more tutorials Try some practice worksheets What's the scientific method?

The scientific method is one way that people can try to find the answer to problems that are bothering them. It's called "scientific", because people like to think of themselves as being very clever, or "scientific" for solving problems. In reality, there's not really anything special about this method, except that it happens to be pretty handy for solving any problem, not just scientific ones.

How do I use the scientific method?


The scientific method is just a list of steps that you need to follow when you're solving a problem. Depending on who you talk to, there are anywhere from five to eight steps in the scientific method. However, all versions of the scientific method involve the person trying to solve the problem experimenting to find an answer. The version of the scientific method that I use in my class has six steps, as follows:

Purpose: You've got a problem that you want to have solved. The purpose step in the scientific method is just a restatement of what you want accomplished. What do you want to find out? What is your goal? You should write just one sentence for your purpose. You'll see what I mean in the upcoming example. Hypothesis: How do you think you can solve the problem? The hypothesis step is always written in the form "If ___________, then ___________. The blank after the "if" is called theindependent variable. The independent variable is just whatever you are going to do to solve the problem. The blank after "then" is the dependent variable. The dependent variable is what you think will happen when you do whatever the independent variable is. For example, if your hypothesis is that "If I take an aspirin, my headache will go away," your independent variable is "taking an aspirin" (this is what you do) and your dependent variable is "the headache will go away" (what happens as a result of your having done something). Materials: What do you need to have in order to see if your hypothesis is true? This part of the scientific method is a list of everything you need to do the experiment. Leave nothing out! Procedure: What are you going to do during this experiment. You should list everything that you are going to do in this section. Even if it seems obvious, write it down. A good rule of thumb: If a six-year-old child can understand what you've written, then you've written it well. If they can't, then you need to go into more detail! Results: When you did the experiment, what happened? What did you see,

hear, smell, etc? You should give a complete accounting of all data that you take (sometimes this is referred to as the "Data" section). There's an old saying among chemists: "If you didn't write it down, then it didn't happen." Make sure you write everything down! Conclusion: What do the results mean? Was your hypothesis correct? This section should be only one sentence long. For example, if you proved the hypothesis that "If I take an aspirin, my headache will go away," then the conclusion should be "I took an aspirin, and my headache went away." Don't make this any longer than it has to be!

An example of the scientific method:


Let's say I have a problem: My car won't start. How would I use the scientific method to solve this problem?

Purpose: I want my car to start Hypothesis: If I put gas in my car, it will start. Materials: 5 gallon gas can, 5 gallons gasoline, money to buy gasoline, a ride to the gas station Procedure: First, I will call my friend Bill and ask for a ride to the gas station. I will take the five gallon gas can and fill it with five gallons of gasoline at the pump. After paying the gas station owner for the gasoline, I will get a ride back to my car and put the gasoline in the tank. Once the gasoline is in the tank, I will attempt to restart the car. Results: The car started on the first try. Conclusion: When I put gas in my car, it started.

It's as simple as that!

Significant Figures
Head back to the helpdesk Find more tutorials Try some practice worksheets

One of the most baffling subjects for students is frequently significant figures. The reason for this is simple: Nobody ever seems to know what theyre supposed to be used for. Why should I care if 100 has one significant figure or 100.0 has four? Fortunately, your friend Mr. Guch is here to help. Lets take a look at the joyous excitement produced by significant figures:

Why do we need significant figures? For some reason, teachers never really tell students why significant figures are important (note to any teachers who are reading this: Im not talking about you. Im talking about those other teachers). Significant figures are important because they tell us how good the data we are using are. (Incidentially, the word data is plural for datum, so even though it seems weird saying that data are [something], its grammatically correct.) For example, lets consider the following three numbers: 100 grams 100. grams 100.00 grams

The first number has only one significant figure (namely, the 1 in the beginning). Because this digit is in the hundreds place, this measurement is only accurate to the nearest 100 grams (i.e. the value of what were measuring is closer to 100 grams than it is to 200 grams or 0 grams). The second number has three significant figures (the decimal makes all three digits significant, as well discuss later). Because the last significant figure is in the ones place, the measurement is accurate to the nearest gram (i.e. the value of what were measuring is closer to 100 grams than it is to 101 grams or 99 grams). The third number has five significant figures (as well talk about later). Because the last significant figure is in the hundredths place, the measurement can be considered to be accurate to the nearest 0.01 grams (i.e. the value of what were measuring is closer to 100.00 grams than it is to 100.01 or 99.99 grams).

In short, when you plug these three numbers into your calculator, theres no difference in how the calculator will manipulate them your calculator neither knows nor cares about how good the numbers its working with are. However, to you, the taker of data, these three numbers tell you whether or not your data is good enough to pay attention to. How do we find the correct number of significant figures? Right now youre thinking to yourself, Mr. Guch, my teacher never mentioned anything you talked about above, but for some reason just likes to ask me a bunch of questions in which I need to figure out how many significant figures a number has. What should I do? What you should do is march right on down to your teacher and tell them that theyve been wasting your time, teaching you something thats totally

irrelevant (because as I mentioned, significant figures are completely irrelevant if you dont understand why theyre important). Of course, your teacher will laugh at you, call your parents, and make you stay after school, so this probably isnt a great idea. What youll probably end up doing is just learning how to figure out the rules for measuring significant figures. However, make sure that you tell your friends why significant figures are handy so they know why theyre bothering with all of this. Rule 1: Any number that isnt zero is significant. Any zero thats between two numbers that arent zeros is significant.

All this means is that if you have actual numbers written down (or zeros between these numbers), they have actual meaning and give you meaningful information. Example: 198, 101, and 987 all have three significant figures.

Rule 2: Any zero thats before all of the nonzero digits is insignificant, NO MATTER WHAT.

Basically, this applies to numbers that are very small decimals. For example, if you have the number 0.000054, there are only two significant figures (the 5 and the 4), because the zeros in front are insignificant. But Mr. Guch, dont those zeros tell me something? Yes and no. The reason that you dont count these numbers as significant is mainly because of rule 4, which well talk about after

Rule 3: Any zero thats after all of the nonzero digits is significant only if you see a decimal point. If you dont actually see a little dot somewhere in the number, these digits are not significant.

Lets consider the numbers 10,000 lbs and 10,000. lbs. The first number is significant only to the nearest ten thousand pounds (only the first 1 is significant) and the second is significant to the nearest pound (all five digits are significant). What the addition of the decimal does is tell us how good our measuring equipment is. The first number, for example, was probably taken by a truck scale (which wouldnt have much use for measuring things to the nearest pound) and the second was probably taken by a bathroom scale (which requires much greater precision).

Rule 4: When you write numbers in scientific notation, only the part before the x is counted in the significant figures. (Example, 2.39 x 104 has three significant figures because we only worry about the 2.39 part).

Lets go back to rule #2, in which we said that 0.000054 had two significant figures. The reason for this is that if we convert it into scientific notation, we end up with the number 5.4 x 10-5. If we said that all the zeros in front were significant in rule 2, then wed have the weird case where the same number had either two significant figures or seven significant figures, depending on

how we write it. Since that kind of confusion isnt too cool, we ignore them in significant figures to make our lives easier.

How about some practice problems? Knock yourself out. Click HERE for a practice worksheet.

Unit Conversions
Head back to the helpdesk Find more tutorials Try some practice worksheets

One of the things that many students have problems with is unit conversions. Unfortunately, unit conversions are really important to know how to do. For example, let's say that somebody wants to do a calculation of the number of moles of water in your body. This seems like it should be easy, but you probably know your weight in pounds (if you live in the US) rather than kilograms - since mole conversions are in grams, you need to be able to convert between pounds and kilograms to do this. Fortunately, there's a way around this. It's called the "factor-label method", or the "t-chart method". Whatever you call it, the idea is the same. The idea behind this method is simple. In every problem, they give you a number you need to convert. We'll refer to this as "what you know", since it's the given in the problem. In every problem you're also given something you need to find. We'll refer to this as "what you want to know". When doing conversions between what you know and what you want to know, you'll need to have some unit conversion factors. These factors are along the lines of "there are 12 inches in one foot" - nothing too disturbing. Let's do an example:

Example: Convert 10 meters to inches. There are 2.54 centimeters in 1 inch.

How to solve:

Go through these steps to make your life much easier: 1) Draw a great big T. It should look like this:

2) Put the number the problem gives you to convert in the top left of this T. Since the problem told you that you've got "10 meters", that's what you should write up there. Check it out:

3) Put the unit of that number in the bottom right part of the T. In this case, the unit is "meters", so you just write that in the bottom right. Have a look:

4) Write the unit of what you want in the top right. Now, in this question we may have a problem. Here it is: Do you know what the conversion factor is between meters and inches? Come on, off the top of your head! Quick, what is it? What do you mean you don't know?

Well, we have a problem then. If you can't convert between meters and inches, I guess you can't do the problem. It's unsolvable! Or is it? Maybe, just maybe we can convert meters to something that can be converted to inches. If we're really lucky, the question might give you a hint as to what you can try. Let's take a look at the question again:
Example: Convert 10 meters to inches.

There are 2.54 centimeters in 1

inch.
I don't know about you, but I just had an idea! Maybe instead of going directly from meters to inches, we could go from meters to centimeters, and then from centimeters to inches! I'm a genius! As a result, we'll put the unit "centimeters" in the top right.

5) Write the unit conversion factor in front of the units from steps 3 and 4. There are three ways to do this:

Sometimes the unit conversion factor is so simple that it doesn't need to be told to you - it's assumed that you know it off the top of your head. Examples of this include the factors "12 inches in 1 foot", "3 feet in 1 yard", "60 minutes in 1 hour", "24 hours in 1 day", and so on. If this is the case for a problem, then all you need to do is look back in your memory to figure it out. When the unit conversion factor isn't simple, the problem will give it to you. For example, in this problem, the problem tells you that there are 2.54 centimeters in 1 inch. That'll probably come in handy, and is obscure enough that you don't need to memorize it. When converting from metric units to other metric units, you put a "1" in front of the unit with the prefix, and write the "multiplier" in front of the unit without the prefix. The common multipliers for metric conversions are as follows: Prefix Multiplier

micro milli centi deci kilo mega giga

0.000001 (10-6) 0.001 (10-3) 0.01 (10-2) 0.1 (10-1) 1,000 (103) 1,000,000 (106) 1,000,000,000 (109)

In this problem, that's what we need to do because we're converting between meters and centimeters. Since "centimeters" has a prefix (centi), we write a 1 in front of it. Since "meters" has no prefix, we put the multiplier for centimeters in front of it. From the chart above, we can see that the multiplier is 0.01. Let's do it!

6) If you're not done with the conversion in one step, draw another vertical line in the t-chart after the first step and start over at step 3. I'll assume that you can follow along with what we're doing at this point, so I'll just show you the steps you need to follow in a bunch of different diagrams. If you have problems following what we're doing, refer to the steps above. Here's what it looks like when you draw the vertical line and stick the unit in the bottom right (from step 3):

Since we want to convert to inches and we have a unit conversion factor between centimeters and inches, write inches in the top right:

Since we know that the unit conversion factor between inches and centimeters is "2.54 centimeters in 1 inch", write "2.54" in front of centimeters and "1" in front of inches:

7) Since you've got the t-chart filled out, all you have left to do is multiply all the numbers on the top together and divide them by the product of all the numbers on the bottom. The unit of this answer will be "inches", since "meters" and "centimeters" cancel out. As a result, we get:

And that's your answer! Yeeeeha!

Graphing guidelines

Head back to the helpdesk Find more tutorials Try some practice worksheets

One of the most common things students have problems with is making good graphs for their experimental data. Fortunately, good ol' Mr. Guch has some suggestions that you can use to get an A+ on every graph you do for a lab or class. I've included an example of a good graph and a bad graph after the rules, so you know what these will look like.

Mr. Guch's rules for Good Graph Making:


1. Always give your graph a title in the following form: "The dependence of (your dependent variable) on (your independent variable). Let's say that you're doing a graph where you're studying the effect of temperature on the speed of a reaction. In this reaction, you're changing the temperature to known values, so the temperature is your independent variable. Because you don't know the speed of the reaction and speed depends on the temperature, the speed of the reaction is your dependent variable. As a result, the title of your graph will be "The dependence of reaction rate on temperature", or something like that.

2. The x-axis of a graph is always your independent variable and the y-axis is the dependent variable. For the graph described above, temperature would be on the x-axis (the one on the bottom of the graph), and the reaction rate would be on the y-axis (the one on the side of the graph)

3. Always label the x and y axes and give units. Putting numbers on the x and y-axes is something that everybody always remembers to do (after all, how could you graph without showing the numbers?). However, people frequently forget to put a label on the axis that describes what those numbers are, and even more frequently forget to say what those units are. For example, if you're going to do a chart which uses temperature as the independent variable, you should write the word "temperature (degrees Celsius)" on that axis so people know what those numbers stand for. Otherwise, people won't know that you're talking about temperature, and even if they do, they might think you're talking about degrees Fahrenheit.

4. Always make a line graph Never, ever make a bar graph when doing science stuff. Bar graphs are good for subjects where you're trying to break down a topic (such as gross national product) into it's parts. When you're doing graphs in science, line graphs are way more handy, because they tell you how one thing changes under the influence of some other variable.

5. Never, EVER, connect the dots on your graph! Hey, if you're working with your little sister on one of those placemats at Denny's, you can connect the dots. When you're working in science, you never, ever connect the dots on a graph. Why? When you do an experiment, you always screw something up. Yeah, you. It's probably not a big mistake, and is frequently not something you have a lot of control over. However, when you do an experiment, many little things go wrong, and these little things add up. As a result, experimental data never makes a nice straight line. Instead, it makes a bunch of dots which kind of wiggle around a graph. This is normal, and will not affect your grade unless your teacher is a Nobel prize winner. However, you can't just pretend that your data is perfect, because it's not. Whenever you have the dots moving around a lot, we say that the data is noisy, because the thing you're looking for has a little bit of interference caused by normal experimental error. To show that you're a clever young scientist, your best bet is to show that you KNOW your data is sometimes lousy. You do this by making a line (or curve) which seems to follow the data as well as possible, without actually connecting the dots. Doing this shows the trend that the data suggests, without depending too much on the noise. As long as your line (or curve) does a pretty good job of following the data, you should be A-OK.

6. Make sure your data is graphed as large as possible in the space you've been given. Let's face it, you don't like looking at little tiny graphs. Your teacher doesn't either. If you make large graphs, you'll find it's easier to see what you're doing, and your teacher will be lots happier.

So, those are the steps you need to follow if you're going to make a good graph in your chemistry class. I've included a couple of examples of good and bad graphs below so you know what these things are supposed to look like.

Examples of Good and Bad Graphs


All those rules I gave you above are true and are handy to know, but it's usually a bad idea to give rules without showing you what they mean. Below are two examples of graphs. One is a bad graph (which you may be guilty of making) and the other is a good graph (which is what I always make).

A bad graph!

Let's see what's wrong with this graph:


There's no title. What's it a graph of? Who knows? There are no labels on the x or y axis. What are those numbers? Who knows? There are no units on the x or y axis. Is this a graph of speed in miles per hour or a graph of temperature in Kelvins? Who can tell? Somebody played "connect the dots". This should be a nice straight line which goes through the points or a curve that tends to follow them.

A good graph!

Doesn't the clarity and beauty of this graph just make you want to cry? Well, maybe that's overstating it a little bit, but it sure does make more sense than the first one, doesn't it? I'm starting to mist up right now.*
*No, I'm not.

Metric Prefixes and Scientific Notation

Head back to the helpdesk Find more tutorials Try some practice worksheets

In this section:

The metric system Metric prefixes How to convert "regular numbers" into scientific notation

Overview: Many of you have trouble with metric prefixes and scientific notation. In this section, we'll start by learning why we have metric prefixes and what the metric prefixes are used for. Toward the end, we'll talk about the wonders and joys of scientific notation. Hold on to your hats! The Metric System

If you already live in a country that uses the metric system (which is practically anywhere but the United States), you can probably skip right past this section to the part about scientific notation. However, if you're an American, chances are you don't really have any idea about how to use the metric system. Don't sweat it - we'll learn about it now.

Why use the metric system? If youve been using nonmetric units your whole life, you may be wondering whats so great about the metric system. After all, youre probably very happy being five feet, ten inches tall and weighing 175 pounds. Its nice knowing that the temperature in your house is 680. Whats the big deal? The problem is not that these units cant be used, because clearly they can. The problem is that its very hard to convert between these units. For example, can you tell me how many inches tall you are? Probably not. However, using the metric system, this problem is very easy.

What are the common metric units? The common metric units are: Seconds: The main unit of time, this is abbreviated as s. Meters: This is the main unit of distance, abbreviated as m. Degrees Celsius: The unit of temperature, abbreviated as 0C. Temperature can also be measured in Kelvins (K), which are found by adding 273 to the temperature in degrees Celsius. Kilograms: The unit of mass, abbreviated as kg. This is kind of a weird one, because the prefix kilo means thousand, so kilograms means thousand grams. Youd think that grams would be the unit of mass, but for some reason, its not. Hertz: The unit of frequency, abbreviated as (Hz). If something happens once a second, it happens with a frequency of 1 hertz. Joule: The unit of work/energy, abbreviated as (J). Pascals: The unit of pressure, abbreviated as (Pa). There are 101,325 pascals (101.325 kPa) in 1 atmospere.

Some handy conversion factors:


seconds to minutes: There are 60 seconds in 1 minute meters to inches: There are 0.0254 meters in 1 inch centimeters to inches: There are 2.54 centimeters in 1 inch kilograms to pounds: There are 2.21 pounds in 1 kilogram pascals to atmospheres: There are 101.325 pascals in 1 atmosphere kilopascals to atmospheres: There are 101.325 kPa in 1 atm. joules to calories: There are 4.184 joules in 1 calorie

Metric Prefixes: The following prefixes are used to make metric prefixes easier to work with. For example, its not too handy to say that you live 345,000 meters from your grandparents. Its much nicer to say that you live 345 kilometers away: Below, for your enjoyment, is a table that will allow you to use the more common prefixes like a star:
Prefix (symbol) nano (n) micro ( ) milli (m) centi (c) deci (d) kilo (k) mega (M) giga (G) What it means 10 10 10 10 10 10 10 10
-9 -6 -3 -2 -1 3 6 9

Example 0.000000001 meters = 1 nm 0.000001 meters = 1 m 0.001 meter = 1 mm 0.01 meter = 1 cm 0.1 meter = 1 dm 1,000 meters = 1 km 1,000,000 meters = 1 Mm 1,000,000,000 meters = 1 Gm

Scientific notation: Scientific notation is another way to express very large or very small numbers so that theyre more understandable. Its easier to say that a speck of dust weighs 1.2 x 10-6 grams than to say it weighs 0.0000012 grams. To convert any number to scientific notation, follow these steps: 1) Convert the number youre converting into a number between 1

and 10 by moving the decimal either to the left or to the right. Examples:

45,000 is converted to 4.5 by moving the decimal four spaces to the left 0.00045 is converted to 4.5 by moving the decimal four spaces to the right

2) Write the number that you came up with in step one, followed by x 10. Examples:

45,000 = 4.5 x 10 0.00045 = 4.5 x 10

3) Recall how many decimal places you moved the decimal point in step one. If the number that youre converting is greater than 10, write a positive number as a superscript above the x 10 from step 2. If the number youre converting is less than one, write a negative number.

Examples:

45,000 = 4.5 x 104 0.00045 = 4.5 x 10-4

And thats all there is to it!

How to calculate the molar masses of chemical compounds


Head back to the helpdesk Find more tutorials Try some practice worksheets
What is molar mass? Molar mass is the weight of one mole (or 6.02 x 1023 molecules) of any chemical compounds. Molar masses of common chemical compounds that you might find in the chemistry laboratory can range between 18 grams/mole for compounds like water to hundreds of grams per mole for more complex chemical compounds. The lightest possible chemical that one can have under normal conditions is hydrogen gas, or H2. There is no limit to how heavy a chemical compound can be - it is not uncommon formacromolecules (large organic or bioorganic compounds such as DNA) to weigh thousands of grams per mole. How can I find the molar mass of an element?

The molar mass of elements is found by looking at the atomic mass of the element on the periodic table. For example, if you want to find the molar mass of carbon, you would find the atomic mass of carbon on the periodic table, and this is equal to the molar mass in grams per mole. So, in our example, carbon has a molar mass of 12.01 grams per mole. There are a few exceptions to this rule. In some cases, the element is usually found in a different form than just one unbonded atom. In the case of hydrogen, nitrogen, oxygen, fluorine, chlorine, bromine, and iodine, the element is diatomic, meaning that each molecule of the element has two atoms of that element stuck together. As a result, the formula of hydrogen is H2, nitrogen is N2, etc. This gets weirder for a couple of cases... phosphorus is normally found in clumps of four atoms, P4, and sulfur is found in clumps of eight atoms, or S8. Still, aside from the exceptions above, all elements have the same molar mass as the atomic masses on the periodic table.

How can I find the molar mass of a chemical compound?

For any chemical compound that's not an element, we need to find the molar mass from the chemical formula. To do this, we need to remember a few rules: 1. Molar masses of chemical compounds are equal to the sums of the molar masses of all the atoms in one molecule of that compound. If we have a chemical compound like NaCl, the molar mass will be equal to the molar mass of one atom of sodium plus the molar mass of one atom of chlorine. If we write this as a calculation, it looks like this: (1 atom x 23 grams/mole Na) + (1 atom x 35.5 grams/mole Cl) = 58.5 grams/mole NaCl

2. If you have a subscript in a chemical formula, then you multiply the number of atoms of anything next to that subscript by the number of the subscript. For most compounds, this is easy. For example, in iron (II) chloride, or FeCl2, you have one atom of iron and two atoms of chlorine. The molar mass will be equal to (1 atom x 56 grams/mole Fe) + (2 atoms x 35.5 grams/mole of chlorine) = 127 grams/mole of iron (II) chloride. For other compounds, this might get a little bit more complicated. For example, take the example of zinc nitrate, or Zn(NO3)2. In this compound, we have one atom of zinc, two atoms of nitrogen (one atom inside the brackets multiplied by the subscript two) and six atoms of oxygen (three atoms in the brackets multiplied by the subscript

two). The molar mass of zinc nitrate will be equal to (1 atom x 65 grams/mole of zinc) + (two atoms x 14 grams/mole of nitrogen) + (six atoms x 16 grams/mole of oxygen) = 189 grams/mole of zinc nitrate. For all other compounds, the general idea is the same. Basically, you should know how to find the molar masses of any chemical compound now. In the next and final section, I'll give you some practice problems, followed by a solution key...

Some sample problems:

These are the kind of molar mass calculation probems I might ask you on a quiz. The solutions are given at the end. Give the molar masses of the following compounds: 1. sodium fluoride 2. potassium hydroxide 3. copper (I) chloride 4. manganese (IV) oxide 5. calcium sulfate 6. magnesium phosphate

Solutions: 1. In sodium fluoride, there is one atom of sodium and one atom of fluorine. The molar mass will then be: (1 atom x 23 grams/mole of sodium) + (1 atom x 19 grams/mole of fluorine) = 42 grams/mole of sodium fluoride 2. In potassium hydroxide, there is one atom of potassium, one atom of hydrogen, and one atom of oxygen. The molar mass will then be (1 x 39 grams) + (1 x 1 gram) + (1 x 16 grams) = 56 grams/mole of potassium hydroxide 3. In copper (I) chloride, there is one atom of copper and one atom of chlorine. The molar mass is then (1 x 63.5 grams) + (1 x 35.5 grams) = 99 grams/mole of copper (I) chloride 4. In manganese (IV) oxide, there is one atom of manganese and two atoms of

oxygen. The molar mass is then (1 x 55 grams) + (2 x 16 grams) = 87 grams/mole of manganese (IV) oxide 5. In calcium sulfate, there is one atom of calcium, one atom of sulfur, and four atoms of oxygen. The molar mass is then (1 x 40 grams) + (1 x 32 grams) + (4 x 16 grams) = 136 grams/mole of calcium sulfate 6. In magnesium phosphate, there are three atoms of magnesium, two atoms of phosphorus, and eight atoms of oxygen. (The formula is Mg3(PO4)2). The molar mass is then (3 x 24 grams) + (2 x 31 grams) + (8 x 16 grams) = 262 grams/mole of magnesium phosphate

All About Mole Calculations


Head back to the helpdesk Find more tutorials Try some practice worksheets

How to do calculations between moles, atoms or molecules, and grams of a substance One of the main problems that beginning chemistry students have is in doing conversions between grams, moles, and molecules (or atoms). Usually, a question will be asked of you in the following form: How many moles are in 22 grams of copper metal? If you're confused by this problem, don't worry. Most people are when they start doing this kind of problem. To make life easier for you, I put together a "road map" which tells you exactly what you need to do to convert between atoms (or molecules), grams, and moles.

You should read this picture the same as you would read a subway map. For example, if you want to go from King Street in Alexandria to National Airport on the Blue line of the Metro, you need to first go to the Braddock Road station. The same thing is true here if a problem tells you to go from atoms to grams, you need to first go through moles before you do anything else. In our example that we're discussing, though, we are making a one-stop trip. Example: How many moles are in 22 grams of copper metal? In all problems like this, you need to go through four steps to find a solution.

The Four Steps to Solving Mole Problems:

Step 1: Figure out how many parts your calculation will have by using the diagram Looking at the diagram above, we can see that we are going between grams and moles, which is a one-step conversion. Furthermore, we can see that we need to use the atomic mass of copper as our conversion factor. Step 2: Make a T-chart, and put whatever information the problem gave you in the top left. After that, put the units of whatever you were given in the bottom right of the T, and the units of what you want to find in the top right.

In this case, the problem gave you "22 grams of copper" as the starting information. Because this is what you were given, put "22 grams of copper" in the top left of the T. Since "grams of copper" is the unit of what you were given, put this in the bottom right of the T. Since you want to find out how many moles of copper are going to be made, put "moles of copper" as your unit in the top right. When you've done this, your calculation should look like this:

Step 3: Put the conversion factors into the T-chart in front of the units on the right. As we saw from the "map", the conversion factor between grams and moles is the atomic mass of copper. Because we measure atomic mass in grams, you need to put the atomic mass in front of the unit "grams of copper". What do you put in front of moles? Whenever you do a calculation of this kind, you need to put "1" in front of moles, like you see here:

Step 4: Cancel out the units from the top left and bottom right, then find the answer by multiplying all the stuff on the top together and dividing it by the stuff on the bottom. In this case, you'd multiply 22 by one and divide the result by 63.5. Your answer, 0.35 moles of copper:

And that's how you do a one-step problem of this kind!

Solving Two-Step Mole Calculation Problems: What happens if we need to solve a problem that requires we not just go from one box in the next in our diagram, but across the entire diagram? Well, it means that we need to do two steps in our calculation. Let's see that "map" again to see what I mean:

If we were asked to convert 22 grams of copper to atoms of copper, we'd have to go from one end of the map to the other. Instead of doing a simple one step calculation, we'd need to do a two-step calculation, with the first step going from grams to moles and the second step going from moles to atoms. How can we solve this kind of problem? Well, we start off by doing the same thing that we did in our last example: We had to convert grams to moles before, and we can see from the map that we have to convert grams to moles now, too. To refresh your memory, here's the calculation from last time:

In the next step, we do the same thing over again, except that we need to add another T to the T-chart. When you do this, take the units of the thing at the new top left and put them on the bottom right (in this case, moles). Then take the units of what you want (in this case, atoms) and put it in the top right. Finally, put in your conversion factors, which from the chart above is Avogadro's number, or 6.02E23. Since this number refers to the number of atoms in a mole of a substance, we put this in front of "atoms of copper". Again, put the number "1" in front of moles, because we're saying that there are 6.02E23 atoms in ONE mole of an element. When we add all these terms in, we can cross out the units that cancel out, as shown. To get the answer, multiply all the numbers on the top together and divide by the numbers on the bottom. Your answer should then be set up like this:

And that's how you do mole problems!

How can I balance an equation?

Head back to the helpdesk Find more tutorials Try some practice worksheets

Why do we need to balance equations, anyway?


Remember that at the beginning of the year, we did a lab where we added baking soda to vinegar and collected the whole mess in a rubber glove? Well, most of you were able to show that the mass of the stuff that we made was the same as the mass of the stuff we started with. (If you weren't one of those lucky people, then let me be the first to tell you this: The mass of the stuff that you make in a chemical reaction is the same as the mass of the stuff that you start with). This is called the Law of Conservation of Mass. Now, this shouldn't really be all that surprising, considering that this is true for most everything else in life. For example, when I make my world-famous chili, the weight of the chili that I make is the same as the weight of all the ingredients added together. As it is with chili, so it is with chemical reactions. Now, when we write chemical equations, we need to have the formulas for the reagents on the left side (the stuff that's going to do the chemical reaction) and the formulas for the products (the stuff you make) on the right. If we were to simply put the formulas of the chemicals on the left and right without saying how much of it was going to react, then we would run the risk of saying that the mass of what we end up with is different than the mass of what we started with. This would be the same thing as writing a recipe where we didn't specify how much of each ingredient is needed to make the chili. The bottom line: You need to balance the equations by sticking numbers in front of the chemicals on the left and right sides of the equation, like it or not. How can you do this? Check out the next section, titled...

... Balancing Chemical Equations


OK. You know why you need to balance chemical equations, but you don't yet know how to do it. It turns out that I'm star who knows how to explain things in a way that even the dumbest people know how to follow. And, hey, if the dumbest people can figure it out, so can you! Listen: There are four easy steps that you need to follow to make this work. Here they are: 1. Get yourself an unbalanced equation. I might give this to you, or I might make you figure it out. Either way, if you don't have an equation with all the chemical formulas and the arrow and all that other stuff, then you're out of luck. 2. Draw boxes around all the chemical formulas. Never, ever, change anything

inside the boxes. Ever. Really. If you do, you're guaranteed to get the answer wrong. 3. Make an element inventory. How are you going to know if the equation is balanced if you don't actually make a list of how many of each atom you have? You won't. You have to make an inventory of how many atoms of each element you have, and then you have to keep it current throughout the whole problem. 4. Write numbers in front of each of the boxes until the inventory for each element is the same both before and after the reaction. Whenever you change a number, make sure to update the inventory - otherwise, you run the risk of balancing it incorrectly. When all the numbers in the inventory balance, then the equation can balance, and you can relax and enjoy a delicious bowl of Mr. Guch's chili. Read on for an example...

An example of equation balancing:


Let's say I ask you the following thing on a test: "Balance the equation that takes place when sodium hydroxide reacts with sulfuric acid to form sodium sulfate and water." How do we solve this using the steps above? 1. Get yourself an unbalanced equation. Here's where you use your knowledge of formulas to help you out. If you know what the formula of sodium hydroxide, sulfuric acid, sodium sulfate, and water are, you'd be able to write the following unbalanced equation:

2. Draw boxes around all the chemical formulas. This is the step that people frequently don't do because they feel that it's a stupid thing to do. Those people are morons. Ignore them. You're drawing those boxes so that you'll be sure not to mess around with the formulas to balance the equation. While they all suffer in the pits of academic hell, you'll be laughing from the honor roll. Here's what the equation looks like:

3. Make an element inventory. In this inventory, your job is to figure out how many atoms of each element you have on the left and right sides of the equation. Now, if you look at the equation, you should be able to see that on the left side of the equation there is one sodium atom, five oxygen atoms (one from the sodium hydroxide, four from the sulfuric acid), three hydrogen atoms (one from the sodium hydroxide, two from the sulfuric acid), and one sulfur atom. On the right side of the equation, there are two atoms of sodium, one atom of sulfur, five atoms of oxygen (four from the sodium sulfate and one from the water), and two atoms of hydrogen. Thus, your element inventory should look like this:

4. Write numbers in front of each of the boxes until the inventory for each element is the same both before and after the reaction. Now, what happens when we put a number in front of a formula? Basically, anything in that box is multiplied by that number, because we're saying that we have that many of that kind of molecule. So, looking at the inventory, what should we do? Well, we can see that on the left side of the inventory, there is one atom of sodium and on the right there are two. The solution: Stick a "2" in front of the sodium hydroxide on the left side of the equation so that the numbers of sodium atoms are the same on both sides of the equation. When we do this, the new atom inventory should look like this: (I'll let you figure out how this is done)

Now what? Well, looking at the new inventory, we can see that we now have two sodium atoms on both the left and the right sides, but the others still don't match up. What to do? You can see from the inventory that on the right side of the equation, there are two hydrogen atoms and on the left there are four. Using your amazing powers of mathematics (and hopefully not needing to use a calculator), you can see that two multiplied by the number two becomes four. That's what you need to do. How? Put a "2" in front of the water on the right side of the equation to make the hydrogens balance out. Now that this is done, you should make a new inventory that looks something like this:

Since both sides of the inventory match, the equation is now balanced! All other equations will balance in exactly the same way, though it might take a few more steps in some cases.

Problems you might encounter:


Is it all as easy as I made it look above? Well, yes and no. Yes, it should work all the time. No, sometimes you need to do some tricks to find the right numbers to add into the equation. For example, what happens when you do the inventory, and you find that there are two atoms of element X on the left side of the equation and three on the right. How can you make those match? When you run into this problem, find the lowest common denominator of those two numbers, and then put the numbers in front of those two boxes which allow the inventory on both sides to match. In the element X example, the lowest common denominator of two and three is six, so you'd put a "3" in front of the molecule on the left, and a "2" in front of the one on the right. Element X will then match up, and you can use a new inventory to see what else needs to be done. Another common problem: What happens when the only way you can get a problem to work out is to make one of the numbers a decimal or fraction? When this happens, find the largest molecule in the equation and stick a "2" in front of it. Then start the problem over. Will this work all the time? Well, no. But it will work sometimes, and give you a new strategy for hard problems. Most importantly: Always remember to keep the inventory of the elements current! If you try to keep it in your head, you'll screw it up. Nobody can keep a bunch of changing numbers in their head for very long. I certainly can't, and you can't either.

Some practice problems:


Here are some practice problems. The solutions are in the section below this one. 1. __NaCl + __BeF2 --> __NaF + __BeCl2 2. __FeCl3 + __Be3(PO4)2 --> __BeCl2 + __FePO4 3. __AgNO3 + __LiOH --> __AgOH + __LiNO3 4. __CH4 + __O2 --> __CO2 + __H2O

5. __Mg + __Mn2O3 --> __MgO + __Mn

Solutions for the practice problems:


1. 2 NaCl + 1 BeF2 --> 2 NaF + 1 BeCl2 2. 2 FeCl3 + 1 Be3(PO4)2 --> 3 BeCl2 + 2 FePO4 3. 1 AgNO3 + 1 LiOH --> 1 AgOH + 1 LiNO3 4. 1 CH4+ 2 O2 --> 1 CO2 + 2 H2O 5. 3 Mg + 1 Mn2O3 --> 3 MgO + 2 Mn

For more practice problems, click here. To go back to the helpdesk, click here. To email Mr. Guch with your observations, questions, or rants, click here.

Writing Word Equations


Head back to the helpdesk Find more tutorials Try some practice worksheets Read this page in Serbo-Croatian (with special thanks to Jovana Milutinovich!)

Ever wonder how to translate a chemical reaction from a written statement in regular English (or Serbo-Croatian) to equation form? Here's how, using the example below:
Example: When calcium hydroxide reacts with hydrochloric acid in water, dissolved calcium chloride and water are formed. This reaction gives off heat.
How to solve a problem like this:

Step 1: Write the unbalanced equation by translating the written names into

chemical formulas In this case, the formulas you need to know are those for calcium hydroxide, hydrochloric acid, calcium chloride, and water. When you translate these into their formulas, you should get the unbalanced equation:

Ca(OH)2 + HCl --> CaCl2 + H2O


If you've forgotten how to write formulas, visit here for more info about writing formulas for ionic compounds and here for writing formulas for covalent compounds.

Step 2: Balance the equation You need to balance the equation to ensure that the chemical reaction follows the law of conservation of mass, which says that you've got to have the same number of atoms of each element on both sides of the equation. I'll assume for the purposes of this activitiy that you know how to balance equations. If you don't, try visiting here for help. For this reaction, the equation, when balanced, looks like this:

Ca(OH)2 + 2 HCl --> CaCl2 + 2 H2O

Step 3: Figure out the states of each of the chemicals in the equation "States" refers to the form in which you can find a chemical. The states you need to worry about are solid, liquid, gas, and aqueous. Solid, liquid, and gas are probably familiar to you, and "aqueous" is just a fancy word for "dissolved in water". The symbol for a solid is (s), liquid is (l), gas is (g), and aqueous is (aq). You need to make sure you write these in the parentheses, and that you write them right after the formulas in the same place and size that you put the subscripts in the formulas. Check out the example below to see what I mean. Now, the big question is this: How can you tell if something is a solid, liquid, gas, or aqueous? Here are some guidelines that might help you:

The equation might tell you. For example, if something is "dissolved in water", you know it's aqueous. If something is a "powder", this indicates that it's a solid. "Vapors" are gases. Some chemicals are so common that you should be able to figure it out. You should know that carbon dioxide is a gas because you learned that you breathe it out of your lungs after you breathe in oxygen. Likewise, you should have a pretty good idea that water is generally a liquid, except at very low temperatures (when it is solid ice) or very high temperatures (as gaseous steam). If you're not explicity told otherwise, assume that ionic compounds are solids.

That's because ionic compounds have very high melting and boiling points, so they usually are. If they're dissolved in water or in some other form, the equation should tell you. If you're not explicity told otherwise, assume that covalent compounds are liquids. This is actually not that great an assumption because there are a lot of exceptions to this rule, but it's better than nothing. Generally, covalent compounds have fairly low melting and boiling points, and many organic compounds are liquids at room temperature. Still, this is just a vague rule of thumb, and won't always work. All metallic elements but mercury are solids. Mercury is a liquid. All nonmetallic elements are solids, except for the following: Bromine is a liquid; The noble gases, chlorine, fluorine, nitrogen, oxygen, and hydrogen are gases.

So, let's take a look at our equation: (In case you forgot what it was, it was

Ca(OH)2 + 2 HCl --> CaCl2 + 2 H2O


Calcium hydroxide is an ionic compound, so we'll assume it's a solid. Hydrochloric acid is a covalent compound, so we'll assume it's a liquid. The equation tells us that calcium chloride is dissolved in water, so it's aqueous. Water is a liquid, because nothing about the statement told us that the reaction took place at anything but room temperature.

Putting all this stuff together, we get the following equation:

Ca(OH)2(s) + 2 HCl(l) --> CaCl2(aq) + 2 H2O(l)

Step 4: Sticking all the other relevant symbols in here somewhere The last thing we need to do is to stick a bunch of other symbols around here to indicate other relevant things about the reaction. These relevant things may include reaction conditions (things you need to do to make the reaction take place) or indications about whether the reaction is exothermic (gives off heat and feels hot) or endothermic (absorbs heat and feels cold). Here's a list of the symbols you may need to use: Symbol What it means

The reaction requires heat or added energy to occur. This symbol is typically written over the arrow. The reaction takes place at the temperature indicated, in this 0 50 C case 50 degrees Celsius. This symbol is written over the arrow The unit "kPa" stands for kilopascals, which is a unit of pressure. This symbol is also written above the arrow and 50 kPa indicates the pressure at which the reaction should take place. Other pressure units are "atmospheres", "Torr", and "mm Hg". other stuff Do whatever the instructions tell you to do. There are actually

around the arrow H symbols written at the end of the equation

quite a few other symbols which are commonly used, but memorizing them all can be tough. These symbols tell you how much energy is absorbed or given off during the reaction. If energy is absorbed, the reaction is endothermic and H has a positive sign. If energy is given off, the reaction is exothermic and H gas a negative sign. If a number is given here, this indicates how exo or endothermic the reaction is. Common units for H are "kJ/mol" or "kcal/mol".

In the reaction we were given, nothing much was said except that the reaction gives off heat, meaning that it's an exothermic reaction. As a result, the only symbol we really need is a H symbol. Since the amount of heat given off wasn't specified, all we can do is say that H is negative. As a result, our equation looks like this:

Ca(OH)2(s) + 2 HCl(l) --> CaCl2(aq) + 2 H2O(l)


And that's all you need to do!

H = -

Writing Word Equations


Head back to the helpdesk Find more tutorials Try some practice worksheets Read this page in Serbo-Croatian (with special thanks to Jovana Milutinovich!)

Ever wonder how to translate a chemical reaction from a written statement in regular English (or Serbo-Croatian) to equation form? Here's how, using the example below:
Example: When calcium hydroxide reacts with hydrochloric acid in water, dissolved calcium chloride and water are formed. This reaction gives off heat.
How to solve a problem like this:

Step 1: Write the unbalanced equation by translating the written names into chemical formulas In this case, the formulas you need to know are those for calcium hydroxide, hydrochloric acid, calcium chloride, and water. When you translate these into their formulas, you should get the unbalanced equation:

Ca(OH)2 + HCl --> CaCl2 + H2O


If you've forgotten how to write formulas, visit here for more info about writing formulas for ionic compounds and here for writing formulas for covalent compounds.

Step 2: Balance the equation You need to balance the equation to ensure that the chemical reaction follows the law of conservation of mass, which says that you've got to have the same number of atoms of each element on both sides of the equation. I'll assume for the purposes of this activitiy that you know how to balance equations. If you don't, try visiting here for help. For this reaction, the equation, when balanced, looks like this:

Ca(OH)2 + 2 HCl --> CaCl2 + 2 H2O

Step 3: Figure out the states of each of the chemicals in the equation "States" refers to the form in which you can find a chemical. The states you need to worry about are solid, liquid, gas, and aqueous. Solid, liquid, and gas are probably familiar to you, and "aqueous" is just a fancy word for "dissolved in water". The symbol for a solid is (s), liquid is (l), gas is (g), and aqueous is (aq). You need to make sure you write these in the parentheses, and that you write them right after the formulas in the same place and size that you put the subscripts in the formulas. Check out the example below to see what I mean. Now, the big question is this: How can you tell if something is a solid, liquid, gas, or aqueous? Here are some guidelines that might help you:

The equation might tell you. For example, if something is "dissolved in water", you know it's aqueous. If something is a "powder", this indicates that it's a solid. "Vapors" are gases. Some chemicals are so common that you should be able to figure it out. You should know that carbon dioxide is a gas because you learned that you breathe it out of your lungs after you breathe in oxygen. Likewise, you should have a pretty good idea that water is generally a liquid, except at very low temperatures (when it is solid ice) or very high temperatures (as gaseous steam). If you're not explicity told otherwise, assume that ionic compounds are solids. That's because ionic compounds have very high melting and boiling points, so they usually are. If they're dissolved in water or in some other form, the equation should tell you. If you're not explicity told otherwise, assume that covalent compounds are liquids. This is actually not that great an assumption because there are a lot of exceptions to this rule, but it's better than nothing. Generally, covalent

compounds have fairly low melting and boiling points, and many organic compounds are liquids at room temperature. Still, this is just a vague rule of thumb, and won't always work. All metallic elements but mercury are solids. Mercury is a liquid. All nonmetallic elements are solids, except for the following: Bromine is a liquid; The noble gases, chlorine, fluorine, nitrogen, oxygen, and hydrogen are gases.

So, let's take a look at our equation: (In case you forgot what it was, it was

Ca(OH)2 + 2 HCl --> CaCl2 + 2 H2O


Calcium hydroxide is an ionic compound, so we'll assume it's a solid. Hydrochloric acid is a covalent compound, so we'll assume it's a liquid. The equation tells us that calcium chloride is dissolved in water, so it's aqueous. Water is a liquid, because nothing about the statement told us that the reaction took place at anything but room temperature.

Putting all this stuff together, we get the following equation:

Ca(OH)2(s) + 2 HCl(l) --> CaCl2(aq) + 2 H2O(l)

Step 4: Sticking all the other relevant symbols in here somewhere The last thing we need to do is to stick a bunch of other symbols around here to indicate other relevant things about the reaction. These relevant things may include reaction conditions (things you need to do to make the reaction take place) or indications about whether the reaction is exothermic (gives off heat and feels hot) or endothermic (absorbs heat and feels cold). Here's a list of the symbols you may need to use: Symbol

What it means

The reaction requires heat or added energy to occur. This symbol is typically written over the arrow. The reaction takes place at the temperature indicated, in this 0 50 C case 50 degrees Celsius. This symbol is written over the arrow The unit "kPa" stands for kilopascals, which is a unit of pressure. This symbol is also written above the arrow and 50 kPa indicates the pressure at which the reaction should take place. Other pressure units are "atmospheres", "Torr", and "mm Hg". other stuff Do whatever the instructions tell you to do. There are actually around quite a few other symbols which are commonly used, but the arrow memorizing them all can be tough. H These symbols tell you how much energy is absorbed or given symbols off during the reaction. If energy is absorbed, the reaction is written at endothermic and H has a positive sign. If energy is given off, the end the reaction is exothermic and H gas a negative sign. If a

of the number is given here, this indicates how exo or endothermic the equation reaction is. Common units for H are "kJ/mol" or "kcal/mol".

In the reaction we were given, nothing much was said except that the reaction gives off heat, meaning that it's an exothermic reaction. As a result, the only symbol we really need is a H symbol. Since the amount of heat given off wasn't specified, all we can do is say that H is negative. As a result, our equation looks like this:

Ca(OH)2(s) + 2 HCl(l) --> CaCl2(aq) + 2 H2O(l)


And that's all you need to do!

H = -

The Atom in History


Head back to the helpdesk Find more tutorials Try some practice worksheets

Most people think that the atom is a very boring thing to study. However, that cant possibly be true after all, lots of people have studied it for thousands of years. It must be exciting! Lets take a look at what well discuss here:

The Greeks talk and talk and talk about atoms. Some new ideas crop up from somewhere or other. Dalton actually says something smart about atoms. Thomsen plays with electricity. Rutherford plays with gold. Rappin Neils Bohr and the planetary model

Lets get started!

The Greeks and their imaginary atom: Even though the Greeks had very little in the way of high technology, they still felt that they could use the power of their brains to figure out what matter was made up of at the smallest levels. As a result, lots of them talked about it a lot. Democritus was one of these guys. He came up with a model of the atom that said:

Atoms are solid and indestructible. Different atoms have different shapes and sizes this is why different materials have different properties.

Because Aristotle disagreed with him and everybody thought that Aristotle was a big hotshot, practically nobody paid attention to Democritus. The moral of the story: Dont mess with Aristotle. Or something like that.

Some ideas that nobody had ever thought of before: For a really long time, nobody really thought that Aristotle was wrong. Eventually, however, with the advance of science, people started to rethink their devotion to the dead Greek guy. Here are some of the discoveries that changed this:

Law of conservation of mass: The amount of stuff you form in a reaction is equal to the amount of stuff you started with. Law of definite composition: Every chemical compound has one and only one chemical formula. For example, no matter what process you use to make water, the formula will always be H2O. Law of multiple proportions: If two elements can combine to form more than one chemical compound, the ratio of the mass of one element that combines with a fixed mass of the other element will be a whole number ratio for the compounds. Since this doesnt make any sense, lets use the example of two compounds where hydrogen reacts with oxygen: H2O and H2O2(hydrogen peroxide). In the first compound, the amount of oxygen thats needed to combine with 2 grams of hydrogen is 16 grams. In the second compound, the amount of oxygen thats needed to combine with 2 grams of hydrogen is 32 grams. Since the ratio of 32/16 works out to a 2:1 ratio, it follows this law. Seems like a lot of words for a simple idea, huh?

Dalton and his not entirely imaginary atom: In the 1800s, some English guy named John Dalton came up with his own idea of what atoms were like. His theory included the following ideas:

Everything is made of atoms (which is true!) All atoms of an element are identical in every way (which is false, because of the existence of isotopes). Atoms of different elements are different (which is true). Atoms cant be broken (which is true for chemical reactions, but not for nuclear ones). Atoms combine in whole number ratios to form compounds (i.e. you cant have half an atom in a compound this isnt really surprising, given his idea that you cant break an atom) this is true. In chemical reactions, atoms are rearranged (this is true).

Overall, people seemed pretty happy with Daltons laws. That is, until his idea of what atoms are like was disproved by

Thomsen and his cathode ray tube: Since every chemistry textbook in the world shows a picture of the cathode ray tube experiment, Im not going to reproduce it I suggest you turn to it, though, since it might help with my explanation. Anyway, one day Thomsen was goofing around the lab with these cathode ray tubes he found somewhere. What he found was that when he connected these big long hollow tubes to batteries, a beam of light would go from one end to another. Since he had a lot of time on his hands, he decided to figure out what the deal was with the light. After all, if there was nothing in the tube to start with, whered the light come from? He figured, it must come from the electrodes since the electrodes were made of atoms, the atoms must somehow be coming apart. Among other things, Thomsen got a magnet and held it near the beam. When he did this, he found that the beam would bend toward the positive side of the magnet and away from the negative side. From this, he figured that the beam must contain very small particles from the atom and that they must have negative charge. This led directly to his plum pudding model of the atom, named after a dessert that nobody can eat without throwing up. Think of a chocolate chip cookie, instead. His idea was that the dough in the chocolate chip cookie made up most of the atom and that it had positive charge. The chips represented the little tiny bits of negative charge that made up the light he was messing around with unlike the dough, they could leave the atom if you gave them a shove (with a battery, for example). For this discovery, Thomsen is forever known. This, despite the fact that his model was almost instantly disproved.

Rutherford and his gold foil: Rutherford was a scientist who liked to play with radioactive stuff. His favorite radioactive thing to play with was alpha particles, which are helium nuclei (they have a mass of 4 amu and a charge of +2). One day, he decided to shoot a bunch of alpha particles at a really thin piece of gold foil. When he did this (again, you can find MUCH better pictures of this in your textbook than I can make), he found that most of the particles went right through the foil, while some of them either passed through or bounced off at irregular angles. Why is this?

His idea was to come up with a model of the atom in which most of the atom is empty space with electrons floating around in it. The protons, however, are all concentrated in the middle of the atom (called the nucleus) according to this model, the positively charged alpha particles would go straight through the atom most of the time and only be deflected on the rare occasions when they passed very close to the tiny nucleus. For this discovery, we will always know Rutherford as the man with the gold foil. And the bad model of the atom, because we now know that this model isnt right, either.

Random interlude: Chadwick and the discovery of the neutron In 1932, James Chadwick discovered the neutron (which has no charge at all) by doing some really complicated experiments that I dont understand even a little bit. Dont worry, though your teacher probably doesnt understand it either (unless theyre a nuclear scientist or something), so if you just remember that Chadwick discovered it, youre probably fine.

Neils Bohr and the planetary model: As tends to be the case with models of the atom, nobody really bought into Rutherfords model for very long because it turned out to be wrong. Neils Bohr was one of the guys that didnt buy it, due to the discovery that when you add energy to an atom, it gives off light that has only a few very particular colors. Since Rutherfords model didnt explain how this could be, Neils (as his buddies called him) came up with a different model (which, as is the case with many things, is much better drawn in your textbook). His idea was that electrons traveled only in certain circular paths around the nucleus, much as the planets circle the sun. When energy is added to the electrons, the electrons jump from their normal orbit (called the ground state orbital) to a higher energy orbital farther from the nucleus (called the excited state orbital). Since the world likes to exist at low energy more than at high energy, the electrons eventually return to their ground state orbitals. When this happens, the energy that they absorbed is given off as light. Since the color of light is very closely related to its energy, you only see very particular colors of light being given off by the very particular energy differences between the ground state and the excited state.

His idea further went on to say that the energies of the orbitals were different for every atom. As a result, the colors of light given off by every element is unique, which allows us to identify them via the magic of spectroscopy (specifically, this phenomenon is called atomic emission spectroscopy). Like all models of the atom, this was overturned in about 15 minutes by a bunch of guys who invented something called quantum mechanics. However, dont feel sorry for the planetary model of the atom it still has a healthy and thriving life in elementary school textbooks (which for some reason refuse to acknowledge the existence of quantum mechanics).
2007 Ian Guch All Rights Reserved

The Periodic Table - A real nifty invention


Head back to the helpdesk Find more tutorials Try some practice worksheets
What's the deal with the periodic table? The periodic table is basically just a list of all the elements that we know about. Because of the way it's set up, we can use it as a reference source for figuring out all kinds of information.

What are the important parts of the periodic table? 1. Metals, nonmetals, and metalloids: The periodic table tells you where the metallic, nonmetallic, and semimetallic metals are. To the right of the periodic table, starting to the left of boron (element #5, B) you should see a line that looks like a staircase. Elements far to the left of this line are metals, elements to the far right of this line are nonmetals, and elements right around the line on either side are semimetals, or metalloids. To review: Metals are conductors of heat and electricity, malleable, ductile, and

generally solid. Nonmetals may be solids, liquids, or gases, and are poor conductors of heat and electricity. When solids, they are brittle, non-lustrous materials. Metalloids are solids at standard conditions, and are semiconductors of electricity, making them handy for use in the electronics field. Metalloids have properties between that of metals and nonmetals, causing them to have the nickname of "semimetals."

2. The families of the periodic table: The periodic table consists of a whole bunch of different families which share similar properties. Families are columns in the periodic table, also referred to as groups.

Alkali metals are group 1. They are highly reactive elements with low melting and boiling points. They are light, soft metals. They tend to form ions with a +1 charge. Alkaline earth metals are group 2. They are also reactive, but less so than the alkali metals. They are light, soft metals, but stronger and denser than the alkali metals. They tend to form ions with a charge of +2. Transition metals are in groups 3-12. They are less reactive than the alkali and alkaline earth metals, but vary greatly among themselves in reactivity. Generally, these elements form cations, but the amount of positive charge these elements have depends on what the metals are reacting with. Lanthanides are the metals in the 4f part of the periodic table. They are generally reactive, shiny metals with various industrial purposes. Like the transition metals, they form cations with varying amounts of charge. Actinides are metals in the 5f part of the periodic table. Most are radioactive and man-made. Uses of these elements are primarily in the generation of nuclear power or in nuclear explosives. Small amounts of elements such as americium are used in smoke detectors. Chalcogens are group 16 in the periodic table. Starting with oxygen, these elements are mostly nonmetallic and somewhat electronegative, forming ions with a -2 charge. Halogens are group 17 in the periodic table. These elements are highly reactive oxidizers, and all form ions with a -1 charge. All are electronegative. All are also extremely dangerous, especially when inhaled. Noble gases are group 18 in the periodic table. They basically don't react with anything because they have a stable octet. They used to be called the inert gases, but it was found a while back that some can form somewhat unstable compounds with halogens and oxygen. Hydrogen is element #1 in the periodic table. It is unlike any other element, and is fairly reactive. Depending on what it reacts with, it can either form a +1 ion (hydronium ion, or "proton") or a -1 ion (hydride ion) - generally, the hydronium ion is easier to form than the hydride ion.

What else can I do with the periodic table?

Boy, am I ever glad you asked! You can find the electron configurations of any element using the periodic table by remembering that the two rows at the far left are S rows, the transition metals in the middle are D rows, the main block elements to the right are P rows, and the lanthanides and actinides are F rows. Likewise, the first S orbital you write down will be the 1s orbital, the first P orbital you write will be the 2p orbital, the first D orbital you write will be the 3d orbital, and the first F orbital you write will be the 4f orbital. If you're not too clear about how this works, take a look at a periodic table that has the electron configurations written out and see if it doesn't make a bit more sense. You can use the periodic table to figure out which elements are most electronegative. (Electronegativity is a measurement of how much elements try to steal electrons from other atoms they're bonded to.) Electronegativity increases as you move from the left side of the table to the right side because elements on the right want to gain electrons to be like the nearest noble gas. Electronegativity decreases as you move down each family because of the shielding effect (inner electrons tend to shove the outer ones away, so the more electrons you have, the less tightly-bound the outer electrons). Atomic radius increases as you move down each family (there are more energy levels) and decreases as you move from left to right (because the nucleus gets more positive charge but the electrons keep the same amount of energy all the way across). For more detailed information about periodic trends, click HERE.

Email Mr. Guch at misterguch@chemfiesta.com

All About Subatomic Particles


Head back to the helpdesk Find more tutorials Try some practice worksheets

How to find the number of protons, neutrons, and electrons that atoms have, as well as finding the number of valence electrons.

Finding Protons, Neutrons, and Electrons of Elements One of the things students frequently have problems with is determining how many protons, neutrons, and electrons an element has, based only on some sparse information given to you on the periodic table or in a problem. This tutorial will give you everything you need to figure it out.

First, let's examine one of the entries in the periodic table:

This is what the entry for helium is on a normal periodic table. The atomic number is found at the top, and is equal to the number of protons that an element has. The reason that all atoms of the element have the same number of protons is that the number of protons determines what element is present - if there were a different number of protons, you would have a different element. For neutral atoms, the atomic number is also equal to the number of electrons present because the number of electrons equals the number of protons in neutral atoms. The atomic symbol is different for every element. As you might guess, the atomic symbol and atomic number are very closely related to one another. After all, any element that has the atomic number "2" will have the atomic symbol "He". As a result, if you don't know the atomic number of an element, you can look up the atomic symbol on the periodic table to find it. As a resuilt, you can determine the number of protons an element has from the atomic symbol if you have a periodic table handy. The atomic mass is also different for every element, and also for isotopes of the same element. The atomic mass is equal to the number of protons that an element has plus the number of neutrons that an element has. This is because protons have a mass of approximately 1 amu (atomic mass unit) and neutrons have a mass of approximately 1 amu. Electrons don't count in this calculation, because their mass is small enough that it can usually be ignored. So, to recap: Atomic number = number of protons (which, for neutral atoms, is equal to the number of electrons). Atomic symbol allows us to find the atomic number because you can just look it up on the periodic table. Atomic mass = number of protons + number of neutrons. For a worksheet of practice problems of this type, click here. Finding the number of valence electrons of an element

This is easier than you might think. First of all, let's go through a couple of terms: Octet rule: The rule that says that all elements want to have the same electron configurations as the nearest noble gas to them. This is because the noble gases are really stable, and all elements want to be stable in the same way. Elements become like the nearest noble gas by gaining or losing electrons, or by sharing electrons with other atoms. Valence electrons: The number of s- and p- electrons in the outermost energy level. You can find the number of valence electrons by counting backwards from the element you're interested in to the last noble gas. However, and this is important for some elements, when you're counting backwards, you want to skip over the d- and fparts of the periodic table. For example, sulfur has 6 valence electrons, as does selenium. Why does selenium have six valence electrons when there are more elements betwen it and the last noble gas? It's because we skip over the entire dblock of elements. Of course, anybody who's read this far is probably curious as to why we need to worry about the octet rule or the number of valence electrons. Well, I'm glad you asked. The octet rule allows us to determine the charge that an element will have when it forms ionic compounds. To find this charge, you'll count from the element that you're interested in to the nearest noble gas. If counting backwards takes you to the nearest noble gas (as in the case of lithium and magnesium), the charge is equal to +[however many elements you need to count across]. For example, since you count backwards one element from lithium to reach the nearest noble gas (which is helium), the charge is +1. If counting forwards takes you to the nearest noble gas (as in the case of oxygen and chlorine), the charge is equal to -[the number of elements you need to count across]. For example, since you count forwards two elements from oxgyen to reach neon, the charge of oxygen when it forms ionic compounds is -2. The number of valence electrons allows us to determine how many electrons an element has that can do covalent bonding, and for this reason is extremely important for nonmetals. For example, when oxygen bonds with two hydrogen atoms, we know that our resulting Lewis structure will need to show eight electrons, as oxygen has six valence electrons (determined by counting backwards to He) and the two hydrogens each have one valence electron. For more info on how to find Lewis structures, click here.

Questions? Comments? Email them to me at misterguch@chemfiesta.com

Important things to know about ionic compounds

Head back to the helpdesk Find more tutorials Try some practice worksheets

What are ionic compounds?


Ionic compounds are basically defined as being compounds where two or more ions are held next to each other by electrical attraction. One of the ions has a positive charge (called a "cation") and the other has a negative charge ("anion"). Cations are usually metal atoms and anions are either nonmetals or polyatomic ions (ions with more than one atom). Think back to grade school: The same thing that makes the positive and negative ends of a magnet stick to each other is what makes cations and anions stick to each other. Usually, when we have ionic compounds, they form large crystals that you can see with the naked eye. Table salt is one example of this - if you look at a crystal of salt, chances are you'll be able to see that it looks like a little cube. This is because salt likes to stack in little cube-shaped blocks. Sometimes when you see a salt, it looks like a powder instead of a cube. This doesn't mean that the salt is not a crystal - it means that the crystals are just so small that you can't see them with the naked eye. If you were to put the powder under a microscope, chances are that you would see little geometric blocks. So, what are the main properties of salts? Well, I'm sure glad you asked...

All ionic compounds form crystals. So far as I know, there are no exceptions to this. Again, salts like to form crystals because when you have a whole bunch of little electrical positive and negative charges all stuck together, they seem to like to bunch into little stacking groups. The arrangement that these ions like to stack into is different, and is referred to as the "unit cell". There are ten or so different general shapes of unit cells. When you get to graduate school, ask me about them. For high school classes, it's really not all that important. Ionic compounds tend to have high melting and boiling points. When I say "high", what I mean is "very, very high." Most of the time, when you work with ionic compounds in a chemistry class, the melting point is hot enough that you can't melt them with a Bunsen burner. So, why are these temperatures so high? Well, it has to do with the way that ionic materials are held together. Remember how we said above that ionic compounds form crystals? These crystals are basically just great big blocks of positive and negative charges all stuck together. To break the positive and negative charges apart, it takes a huge amount of energy. This means that if we heat up the compound to add energy, it takes a huge amount of energy to break it apart. Ionic compounds are very hard and very brittle. Again, this is because of the way that they're held together. Above, we said that it takes a lot of energy

to break the positive and negative charges apart from each other. This is the reason that ionic compounds are so hard - they simply don't want to move around much, so they don't bend at all. This also explains the brittleness of ionic compounds. It takes a lot of energy to pull ionic charges apart from each other. However, if we give a big crystal a strong enough whack with a hammer, we usually end up using so much energy to break the crystal that the crystal doesn't break in just one spot, but in a whole bunch of places. Instead of a clean break, it shatters. Ionic compounds conduct electricity when they dissolve in water. If we take a salt and dissolve it in water, the water molecules pull the positive and negative ions apart from each other. (This is because of the unusual properties of water, but that's a different story for a different time). Instead of the ions being right next to each other, they wander all over the water. Now, think back to what electricity is - hopefully, you remember that electricity is just the movement of electrons through metals (or anything else). Now, electrons are just negatively-charged particles, and metals have the property that they're good at letting them wander around. Dissolved salts are the same way. When you dissolve the salt in the water, the positive and negative ions in the water allow electrons to flow much better than if you just had water by itself. Voila! Salt water conducts! A question you might have is "Does electricity travel through salt crystals?" Nope. It doesn't. Because the ions are stuck in one place due to the structure of the crystal, the electricity doesn't move around very well. Another good question: "Does water without salt in it conduct electricity?" The answer: Not very well. Water by itself is a lousy conductor. The reason that boneheads who put the hairdryer in the bathtub with them turn into human fritters is that when they wash themselves, all the crud on them gets dissolved in the water. Some of the crud is ionic, so when the dryer hits the water, they get zapped. An interesting "thought experiment" would be to wash all the salt off yourself and then drop a hairdryer in the bathtub with you. In theory, you would be fine. In real life, you'd still become a crispy critter because tap water by itself has ionic compounds dissolved in it anyway. Thus, my warning: If you put a hairdryer in the bathtub with you under ANY conditions, you will fry yourself!

How do we name ionic compounds?


For practice problems with complete solutions, click here. Most ionic compounds (and any I would ever give on a test) have two word names. The first word in the name is the name of the cation, and the second word is the name of the anion. There is no exception to this rule. The best way to go about naming ionic compounds is to take a look at the formula and figure out the names of the cation and anion. When you've got that, just stick them together and you've got the name of the compound. So, how do we name cations? If the cation is a main block element, the name of the cation will just be the name of the element. So, the Na+ ion is the "sodium" ion. Not

too challenging. However, if the cation is a transition metal, what you need to do is to check out whether or not there is more than one possible charge for that element. For example, iron can have a charge of either +2 or +3. As a result, you need to specify whether the cation has a +2 or a +3 charge. When you've done this, just put the number after the name of the element in Roman numerals. For example, the Fe+3 ion just has the name "iron (III)". How about anions? If the anion has only one atom in it, then the name of the anion is the same as the name of the element EXCEPT the end of the element name is taken off and "-ide" is added to the end. Thus, oxygen becomes "oxide", sulfur becomes "sulfide", phosphorus is "phosphide", et cetera. If the anion has more than one atom, then we'd say that it's a "polyatomic ion", meaning (not surprisingly) that the anion has more than one atom. Look up the polyatomic ion in a table (or pull it out of your... uh... memory), and you've got the name. Thus, OH- is "hydroxide", SO42- is sulfate, et cetera.

Handy methods for naming compounds

Naming ionic compounds if you're given the formula


Let's go through this using an example: Fe2(SO4)3 Step One: Name the cation and anion The cation is always the first thing you see in the name, and the anion is always the second thing. In this case, you should recognize that Fe is "iron", and that SO4 is the "sulfate" ion. Generally, if one of these ions has more than one atom in it, you'll need to look it up in a chart. If you're one of my students, you need to have eight of the polyatomic ions memorized: hydroxide, nitrate, nitrite, sulfate, sulfite, carbonate, phosphate, ammonium. If you don't know the formulas, look 'em up. Step Two: Figure out if you need a Roman numeral in the name. If the cation in the compound you're naming is not a transition metal, then you definitely don't need to use a Roman numeral and the naming is done. If there is, then you need to figure out whether or not the cation can exist in more than one charge. If not, then you don't need a Roman numeral. If so, then move on to Step Three... Step Three: Figure out what the Roman numeral should be Basically, this should be fairly easy. A good rule of thumb is that usually the number of anions you have in the molecule is equal to the charge on the cation, and that the number of cations you have is equal to the number of anions. Using our example, there are three sulfate ions, meaning that iron has a charge of +3. Likewise, since there are two iron atoms, the sulfate has a charge of -2. Since iron has a charge of +3 in this compound, the name in this example is iron (III) sulfate.

Step Four: Check your work Look at the answer from the last step, and ask yourself whether the charges are OK. Is +3 a charge that iron can have? Is -2 the charge of the sulfate ion? In this case, the answer to both questions is "yes", so we're finished, and the answer of iron (III) sulfate stands. But... what if we find a mistake when we check our work? In this case, you have to find another way to solve the problem. Take the example of FeS. If we solve the problem using the first three steps, we find that the formula should be iron (I) sulfide. However, if we check this work as we should in step four, we find that iron cannot have a charge of +1, only +2 or +3, and sulfur can only have a charge of -2. In a case like this, you need to find another way to solve the problem. When this happens, look at the anion. In our example of FeS, the anion is the sulfide anion, S-2. If we have one sulfide ion, this means that the total negative charge in the molecule is -2. As a result, iron must have a charge of +2 to counterbalance the -2 charge of sulfur. Since iron has a charge of +2, the name of the compound is iron (II) sulfide.

Giving the formula of an ionic compound if you're given the name


We'll use an example to find the formula of an ionic compound: copper (II) fluoride Step One: Translate the name into the ions In copper (II) fluoride, the cation is the copper (II) ion and the anion is the fluoride ion. Hopefully, you realize that the copper (II) ion is simply Cu2+ and the fluoride ion is F-. If not, then you need to go back and review the rules for naming ions above. Step Two: Put brackets around the ions, but leave the charges on the outside. In this case, the copper (II) ion would be [Cu]2+ and the fluoride ion would be [F]1 . Never change anything in these brackets, ever! Step Three: Put the ions next to each other. When we do this here, we get [Cu]2+[F]-1 Step Four: Cross the charges: The charge on the cation will be equal to the number of anions you have, and the charge on the anion will be equal to the number of cations you have. In our example, you should realize that we have one copper atom (because the charge on fluorine is -1) and two fluoride ions (because the charge on copper is +2). This gives us a formula of: [Cu][F]2 Step Five: Take the brackets away. The final formula for copper (II) fluoride is then CuF2

Exceptions: IF the charges on the ions can be divided by the same number, then do it before you do step four. For example, if you were to find the formula for manganese (IV) oxide, you'd realize in step three that both manganese (IV) and oxygen have charges that can be divided by two. Instead of crossing the +4 for manganese and the -2 for oxygen, you'd simplify it so that you cross a +2 for manganese and a -1 for oxygen. IF we have a polyatomic ion, such as sulfate or ammonium, you need to replace the brackets with parentheses in step five. For example, if you end up with [NH 4]2O as the formula for ammonium oxide at the end of step four, you'd simply replace the brackets with parentheses in step five to give you (NH4)2

Other stuff I might have forgotten above


In no particular order, here's some other stuff about ionic compounds that you might have wondered about:

Ionic compounds are usually formed when metal cations bond with nonmetal anions. The only common exception I know to this is when ammonium is the cation - there's no metal in ammonium, but it forms ionic compounds anyhow. Ions are atoms that have satisfied the octet rule (which for those of you who've been sleeping the last couple of months, states that every atom wants to have eight valence electrons, just like the nearest noble gas). If you have two neutral elements, and one wants to gain electrons to be like the nearest noble gas and the other wants to lose electrons to be like the nearest noble gas, chances are that they will react with each other and make an ionic compound.

Comments, questions, or gripes? Email me at misterguch@chemfiesta.com

All About Covalent Compounds


Head back to the helpdesk Find more tutorials Try some practice worksheets

A voyage into the wonderful world of covalent compounds! Find out how covalent bonding is different from ionic bonding, how the properties of covalent compounds differ from ionic compounds, and how to name covalent compounds. If you're interested in finding out how to draw Lewis structures, that's not on this page, but you can find out how to do it by clicking here.

What's a covalent compound? I'm sure glad you asked! A covalent compound is a compound in which the atoms that are bonded share electrons rather than transfer electrons from one to the other. While ionic compounds are usually formed when metals bond to nonmetals, covalent compounds are formed when two nonmetals bond to each other. The big question that students frequently have is, "Why do elements share electrons? After all, wouldn't electrons rather grab electrons outright? That's what happens when ionic compounds are formed." The reason that nonmetals have to share electrons with each other has to do with electronegativity. Recall that electronegativity is a measure of how much an element pulls electrons away from other elements it is bonded to. Metals generally have very low electronegativities (they don't much want to grab electrons) while nonmetals have high electronegativities (they really want to grab electrons). The reason for this trend is the octet rule, which says that all elements want to have the same number of electrons as the nearest noble gas, because noble gases are unusually stable. When metals bond to nonmetals, ionic compounds are formed because the metal atoms don't want electrons and easily give them to nonmetals that do want electrons. It's a different story when two nonmetals bond with each other. Instead of having one element give electrons to another, we run into a case where we have two elements that have roughly the sameelectronegativity. As a result, neither element can steal electrons from the other. As a result, if either of them are going to be like the nearest noble gas, they'll have to share electrons rather than transfer them. What are the properties of covalent compounds? Covalent compounds have the following properties (keep in mind that these are only general properties, and that there are always exceptions to every rule): 1) Covalent compounds generally have much lower melting and boiling points than ionic compounds. As you may recall, ionic compounds have very high melting and boiling points because it takes a lot of energy for all of the + and - charges which make up the crystal to get pulled apart from each other. Essentially, when we have an ionic compound, we need to break all of the ionic bonds in order to make it melt. On the other hand, when we have covalent compounds we don't need to break any bonds at all. This is because covalent compounds form distinct molecules, in which the atoms are bound tightly to one another. Unlike in ionic compounds, these molecules don't interact with each other much (except through relatively weak forces called "intermolecular forces"), making them very easy to pull apart from each other. Since they're easy to separate, covalent compounds have low melting and boiling points.

2) Covalent compounds are soft and squishy (compared to ionic compounds, anyway). The reason for this is similar to the reason that covalent compounds have low melting and boiling points. When you hit an ionic compound with something, it feels very hard. The reason for this is that all of the ionic bonds which hold together the crystal tend to make it very inflexible and hard to move. On the other hand, covalent compounds have these molecules which can very easily move around each other, because there are no bonds between them. As a result, covalent compounds are frequently flexible rather than hard. Think of it like this: Ionic compounds are like giant Lego sculptures. If you hit a Lego sculpture with your fist, it feels hard because all of the Legos are stuck very tightly to one another. Covalent compounds are more like those plastic ball pits they have at fast food playgrounds for little kids. While the balls themselves are held together very tightly (just like covalent molecules are held together tightly), the balls aren't really stuck to each other at all. As a result, when little kids jump into the ball pits they sink in rather than bouncing off. 3) Covalent compounds tend to be more flammable than ionic compounds. The main reason that things burn is because they contain carbon and hydrogen atoms that can react to form carbon dioxide and water when heated with oxygen gas (that's the definition of a combustion reaction). Since carbon and hydrogen have very similar electronegativities, they are mostly found together in covalent compounds. As a result, more covalent compounds than ionic compounds are flammable. There are a couple of exceptions to this rule. The first is with covalent compounds that contain neither carbon nor hydrogen. These tend not to burn, and if they do, they burn by mechanisms other than the classic combustion reaction. The other exception comes with ionic compounds referred to as "organic salts". These organic salts are ionic compounds in which the anion is basically a big covalent molecule containing carbon and hydrogen with just a very small ionic section. As a result, they burn even though they're technically ionic compounds. 4) Covalent compounds don't conduct electricity in water. Electricity is conducted in water from the movement of ions from one place to the other. These ions are the charge carriers which allow water to conduct electricity. Since there are no ions in a covalent compound, they don't conduct electricity in water. 5) Covalent compounds aren't usually very soluble in water. There's a saying that, "Like dissolves like". This means that compounds tend to dissolve in other compounds that have similar properties (particularly polarity). Since water is a polar solvent and most covalent compounds are fairly nonpolar, many covalent compounds don't dissolve in water. Of course, this is a generalization and

not set in stone - there are many covalent compounds that dissolve quite well in water. However, the majority of them don't because of this rule. Naming Covalent Compounds Covalent compounds are much easier to name than ionic compounds. Here's how you do it: All covalent compounds have two word names. The first word typically corresponds to the first element in the formula and the second corresponds to the second element in the formula except that "-ide" is substituted for the end. As a result, HF is named "hydrogen fluoride", because hydrogen is the first element and fluorine is the second element. If there is more than one atom of an element in a molecule, we need to add prefixes to these words to tell us how many are present. Here are the prefixes you'll need to remember: Number of atoms 1 2 3 4 5 6 7 8 Let's see how this works: Examples: P2O5 - this is named diphosphorus pentoxide, because there are two phosphorus atoms and five oxygen atoms. CO - this is carbon monoxide (you need the "mono-" because there's only one oxygen atom). CF4 - this is carbon tetrafluoride, because there's one carbon atom and four fluorine atoms. Prefix mono- (use only for oxygen) ditritetrapentahexaheptaocta-

Some important exceptions to this naming scheme occur because the compounds were originally named before the methodical naming scheme above became widespread. Nowadays, these names are so common that they're officially

recognized: H2O is "water" NH3 is "ammonia" CH4 is "methane" There are lots of other names for covalent compounds that are commonly used, particularly for organic molecules and acids. However, I'm not going to cover those because they take a really long time to type out, and my fingers are getting tired.

For practice problems involving naming compounds, click here.

Questions, Comments, or Irate Tirades? Email them to me at misterguch@chemfiesta.com

A Sure-Fire Way to Draw Lewis Structures!


Head back to the helpdesk Find more tutorials Try some practice worksheets

Lewis structures are a way to write chemical compounds where all the atoms and electrons are shown. Sometimes, people have a lot of trouble learning how to do this. However, using the methods on this page, you should have very little trouble. The first method given allows you to draw Lewis structures for molecules with no charged atoms, while the second allows you to do it for charged molecules (such as polyatomic ions). How to draw Lewis structures for molecules that contain no charged atoms 1) Count the total valence electrons for the molecule: To do this, find the number of valence electrons for each atom in the molecule, and add them up.

2) Figure out how many octet electrons the molecule should have, using the octet rule: The octet rule tells us that all atoms want eight valence electrons (except for hydrogen, which wants only two), so they can be like the nearest noble gas. Use the octet rule to figure out how many electrons each atom in the molecule should have, and add them up. The only weird element is boron - it wants six electrons. 3) Subtract the valence electrons from octet electrons: Or, in other words, subtract the number you found in #1 above from the number you found in #2 above. The answer you get will be equal to the number of bonding electrons in the molecule. 4) Divide the number of bonding electrons by two: Remember, because every bond has two electrons, the number of bonds in the molecule will be equal to the number of bonding electrons divided by two. 5) Draw an arrangement of the atoms for the molecule that contains the number of bonds you found in #4 above: Some handy rules to remember are these:
o o o o

Hydrogen and the halogens bond once. The family oxygen is in bonds twice. The family nitrogen is in bonds three times. So does boron. The family carbon is in bonds four times.

A good thing to do is to bond all the atoms together by single bonds, and then add the multiple bonds until the rules above are followed. 6) Find the number of lone pair (nonbonding) electrons by subtracting the bonding electrons (#3 above) from the valence electrons (#1 above). Arrange these around the atoms until all of them satisfy the octet rule: Remember, ALL elements EXCEPT hydrogen want eight electrons around them, total. Hydrogen only wants two electrons.

Let's do an example: CO2


Note: Each of the numbers below correspond to the same numbered step above.

1) The number of valence electrons is 16. (Carbon has four electrons, and each of the oxygens have six, for a total of 4 + 12 = 16 electrons). 2) The number of octet electrons is equal to 24. (Carbon wants eight electrons, and each of the oxygens want eight electrons, for a total of 8+16 = 24 electrons). 3) The number of bonding electrons is equal to the octet electrons minus the valence electrons, or 8.

4) The number of bonds is equal to the number of bonding electrons divided by two, because there are two electrons per bond. As a result, in CO2, the number of bonds is equal to 4. (Because 8/2 is 4). 5) If we arrange the molecule so that the atoms are held together by four bonds, we find that the only way to do it so that we get the following pattern: O=C=O, where carbon is double-bonded toboth oxygen atoms. 6) The number of nonbonding electrons is equal to the number of valence electrons (from #1) minus the number of bonding electrons (from #3), which in our case equals 16 - 8, or 8. Looking at our structure, we see that carbon already has eight electrons around it. Each oxygen, though, only has four electrons around it. To complete the picture, each oxygen needs to have two sets of nonbonding electrons, as in this Lewis structure:

How to draw Lewis structures for molecules that contain one or more charged atoms This method is basically the same one you learned above, except that there are a few extra rules to keep track of. Changes in the procedure above are outlined in red for your convenience. 1) Count the total valence electrons for the molecule: To do this, find the number of valence electrons for each atom in the molecule, and add them up. For polyatomic anions, add the charge of the ion to the number of valence electrons. For polyatomic cations, subtract the charge of the ion from the number of valence electrons. 2) Figure out how many octet electrons the molecule should have, using the octet rule: The octet rule tells us that all atoms (including boron) want eight valence electrons (except for hydrogen, which wants only two), so they can be like the nearest noble gas. Use the octet rule to figure out how many electrons each atom in the molecule should have, and add them up. 3) Subtract the valence electrons from octet electrons: Or, in other words, subtract the number you found in #1 above from the number you found in #2 above. The answer you get will be equal to the number of bonding electrons in the molecule.

4) Divide the number of bonding electrons by two: Remember, because every bond has two electrons, the number of bonds in the molecule will be equal to the number of bonding electrons divided by two. 5) Draw an arrangement of the atoms for the molecule that contains the number of bonds you found in #4 above: Some handy rules to remember are these:
o o o o o

Hydrogen and the halogens bond once. The family oxygen is in bonds one, two, or three times. The family nitrogen is in bonds two, three, or four times Boron usually bonds four times. The family carbon is in bonds four times.

A good thing to do is to bond all the atoms together by single bonds, and then add the multiple bonds until the rules above are followed. 6) Find the number of lone pair (nonbonding) electrons by subtracting the bonding electrons (#3 above) from the valence electrons (#1 above). Arrange these around the atoms until all of them satisfy the octet rule: Remember, ALL elements EXCEPT hydrogen want eight electrons around them, total. Hydrogen only wants two electrons. 7) To find the charge on each atom, compare the number of electrons that each atom has to the number of valence electrons it usually has. For this purpose, each bond counts as one electron and each lone pair counts as two electrons. For example, in CO2 above, carbon has four electrons (because it has four bonds) and oxygen has six (two bonds + 4 lone pair electrons). If the number of electrons that the atom has is more than the normal number of valence electrons, the atom has a negative charge. If the number is less than the normal number of valence electrons, the atom has a positive charge. If it's the same, the atom has no charge at all.

Finding Bond Angles, Shapes, and Hybridizations


Gettin' funky, VSEPR style Head back to the helpdesk Find more tutorials Try some practice worksheets

Sometimes people have a hard time with the whole VSEPR thing. In this helpdesk section we'll discuss what VSEPR means, what it's all about, and how you can use a great big flow chart to figure out the bond angles, shapes, and hybridizations of various covalent compounds.

This whole thing assumes, by the way, that you know how to draw Lewis structures. If you don't, click here.

What is VSEPR?
VSEPR stands for Valence Shell Electron Pair Repulsion. It's a complicated acronym, but it means something that's not difficult to understand. Basically, the idea is that covalent bonds and lone pair electrons like to stay as far apart from each other as possible under all conditions. This is because covalent bonds consist of electrons, and electrons don't like to hang around next to each other much because they have the same charge. This VSEPR thing explains why molecules have their shapes. If carbon has four atoms stuck to it (as in methane), these four atoms want to get as far away from each other as they can. This isn't because the atoms necessarily hate each other, it's because the electrons in the bonds hate each other. That's the idea behind VSEPR.

What is hybridization?
Now, one problem with the whole VSEPR thing is that if you have four things stuck to carbon, for example, there are no orbitals that want to get 109.5 degrees apart from each other (109.5 degrees corresponds to the geometric maximum distance the atoms can get apart). After all, s-orbitals go in a complete sphere (360 degrees) and p-orbitals are 90 degrees apart. What happens instead of using s- or p- orbitals is that when covalent bonds are formed, the s- and p- orbitals mix to form something called hybrid orbitals. "Hybrid" just means "mixture of two different things", and that's exactly what a hybrid orbital is. When three p-orbitals with 90 degree angles combine with one s-orbital with 360 degrees, they average to form four sp3 orbitals with 109.5 degree bond angles. Depending on the numbers of s- and p-orbitals that mix, you can get a bunch of different bond angles.

Common shapes you should know


There are a whole bunch of common shapes you need to know to accurately think of covalent molecules. Here they are:

Tetrahedral: Tetrahedral molecules look like pyramids with four faces. Each point on the pyramid corresponds to an atom that's attached to the central atom. Bond angles are 109.5 degrees. Trigonal pyramidal: It's like a tetrahedral molecule, except flatter. It looks kind of like a squished pyramid because one of the atoms in the pyramid is replaced with a lone pair. Bond angles are 107.5 degrees (it's less than tetrahedral molecules because the lone pair shoves the other atoms closer to each other). Trigonal planar: It looks like the hood ornament of a Mercedes automobile,

or like a peace sign with that bottom-most line gone. The bond angles are 120 degrees. Bent: They look, well, bent. Bond angles can be either 118 degrees for molcules with one lone pair or 104.5 degrees for molecules with two lone pairs. Linear: The atoms in the molecule are in a straight line. This can be either because there are only two atoms in the molecule (in which case there is no bond angle, as there need to be three atoms to get a bond angle) or because the three atoms are lined up in a straight line (corresponding to a 180 degree bond angle). There are other types, but we won't worry about them.

Using a flow chart to figure out what hybridization, shape, and bond angle an atom has
Take a look at this flow chart. I'll explain how to use it to find all the stuff above at the end.

Complicated, huh? Here's how to use it: 1) Draw the Lewis structure for the molecule. This vital if you're going to get the answer right. 2) Count the number of "things" on the atom you're interested in. Let's say that you're looking at methane, CH4. If you want to find the bond angles, shape, and

hybridization for carbon, count the number of things that are stuck to it. Now, the vague term "things" refers to atoms and lone pairs. IT DOES NOT REFER TO THE NUMBER OF BONDS! When you look at methane, there are four atoms stuck to it, so you'd go down the line that says "four" toward the green boxes on this chart. People get confused with multiple bonds. Take carbon dioxide, for example. There are four bonds (carbon is double-bonded to each oxygen) but only two oxygen atoms bonded to carbon. In this case, we count two things stuck to carbon, because we only count the atoms, NOT the number of bonds. Likewise, with ammonia there are four things. Three of the things on nitrogen are hydrogen atoms and the fourth is a lone pair. For the purposes of VSEPR, lone pairs count exactly the same as atoms, because they consist of negative charge, too. 3) Count the number of lone pairs that are on the atom you're interested in. IMPORTANT: This does NOT mean to count the number of lone pairs on all of the atoms in the molecule. Lone pairs on other atoms aren't important - what's important is only what's directly stuck to the atom you're interested in. We mentioned above that methane has four things stuck to it. Since all four things are hydrogen atoms, we moved toward the green boxes on the flow chart. When we get to our second question, we find that there are no lone pairs on carbon, so our answer is zero. When we go down the line that says "zero" from that box, we find that methane is sp3 hybridized, with a 109.5 degree bond angle and tetrahedral shape. And, hey, that's what we were looking for!

Some sample problems:


What are the shapes, bond angles, and hybridizations of the following molecules? Use the flow chart and instructions above to figure it out. 1) 2) 3) 4) 5) 6) 7) 8) carbon tetrabromide phosphorus trichloride oxygen the chlorine atom in hydrochloric acid (HCl) boron trichloride CH2O sulfur difluoride either carbon atom in C2H2

The answers are below:

1) 2) 3) 4) 5) 6) 7) 8)

sp3, tetrahedral, 109.5 degrees. sp3, trigonal pyramidal, 107.5 degrees. sp2, linear, no bond angle sp3, linear, no bond angle sp2, trigonal planar, 120 degrees sp2, trigonal planar, 120 degrees sp3, bent, 104.5 degrees sp, linear, 180 degrees

Questions? Comments? Maps to lost treasure? Send them all to me at misterguch@chemfiesta.com!

Types of Solids
Head back to the helpdesk Find more tutorials Try some practice worksheets

An overview of the different types of solids, how they differ from each other, and where you can find them. Where appropriate, Ive put handydandy links to other parts of the site that might help you out. Whats a solid?

Solid refers to the state of matter in which the particles are locked into place without crystal lattices or just kind of stuck together with intermolecular forces so tightly that t on the type of solid were talking about.

Solids differ from liquids in that the particles in liquids, while still stuck together, do ha that gas molecules really dont interact with each other much, flying all over the place What are the different types of solids?

Ionic solids: Ionic solids are solids in which anions and cations (negatively and posi stick together via electrostatic attraction. By electrostatic attraction, I basically me other. When they do this, they form great big crystals in which each ion is surrounde (as theyre called) are really stable, requiring lots of energy (called the lattice energy chloride (table salt) and Epsom salts (magnesium sulfate heptahydrate). For more in

Metallic solids: Metallic solids are solids in which the positively charged nuclei are h bind the whole mess together. These electrons are referred to as being delocalized

in covalent bonds instead, they travel throughout the solid. This allows the atoms in move around with them. A phrase commonly used to describe metallic bonding is el nuclei as floating around in an ocean of negative electrons that hold them together. J kind of bonding.

Network atomic solids: Network atomic solids are great big crystals in which all of t bonds. Because the atoms are all locked into place, these solids usually have proper and boiling point, hard, brittle, and so forth) with the exception that they dont conduct as amethyst, diamond, and ruby) are network atomic solids.

Molecular solids: Molecular solids occur when covalent molecules are held togethe occurs in ice, the intermolecular forces between the molecules are strong enough to k types of solids have much lower melting and boiling points than metallic, network atom holding the molecules together are much weaker than those of the bonds in the other mostly with the properties and behavior of this type of solid.

Atomic solids: Atomic solids occur when noble gases are cooled to really low tempe weakLondon dispersion forces. You wont run into these solids in the real world, be see them.

Amorphous solids: Amorphous solids, unlike the rest of these solids, have no partic particles are just kind of stuck all over the place, with no regular bonding pattern. Som plastic and rubber) because they consist of long molecules which are just kind of tang molecular solids (called glassy solids) are a lot more like network atomic solids becau fashion using covalent bonds. It wont be much of a surprise to find that glass is an e

Questions? Comments? Email them to me at misterguch@chemfiesta.com

All About Intermolecular Forces!


Head back to the helpdesk Find more tutorials Try some practice worksheets

In this helpdesk section we'll answer the following questions:


What the heck is an intermolecular force, anyway? What is each type of intermolecular force and how do they work?

How can we tell the type of intermolecular force holds molecules of a compound to each other? What difference does all this intermolecular force stuff make, anyway?

What the heck is an intermolecular force, anyway?


There are two main types of forces that can occur in a compound, intramolecular forces and intermolecular forces: Intramolecular forces are the forces that keep the atoms in a compound stuck to each other - in other words, they're just chemical bonds. For example, if you're looking at a water molecule, the force that holds the hydrogen atoms to the oxygen are intramolecular forces (covalent bonds). Same deal with sodium chloride - the forces that hold the sodium ions to the chloride ions are intramolecular forces (covalent bonds). Intermolecular forces, on the other hand, are the forces that hold two covalent molecules to one another. For example, if we talk about water, the forces that attract one water molecule to another nearby water molecule are intermolecular forces. These aren't chemical bonds like you've talked about before - instead, they're all based on magnetic attractions between different molecules. This attraction is caused by polarity.

An aside: What's polarity? Polarity is when a molecule acts like a little magnet. This is because one side of the molecule has more electrons than the other side. An example would be HF: The fluorine is very electronegative, so the electrons that would normally be shared equally between the hydrogen and the fluorine tend to spend more time near the fluorine atom, causing the fluorine atom to have a partial negative charge and the hydrogen atom to have a partial positive charge. Because these partial charges are present in the molecule, it behaves like a little magnet.

What the heck is each type of intermolecular force and how do they work, anyway?
There are three main types of intermolecular forces. I'll discuss each of them below, starting with the easiest one: 1) Dipole-dipole forces: In the "aside" above, we discussed polarity. Dipole-dipole forces are simply when two molecules that both behave like little magnets stick to each other. The strength of the dipole-dipole force depends on how polar the molecule is, because the more polar each molecule is, the stronger the magnetic force present. The molecules of every polar compound are stuck to each other by these dipole-dipole forces.

2) Hydrogen bonds: Hydrogen bonds also take place in polar molecules. However, the molecules that undergo hydrogen bonding all contain a hydrogen atom that's bonded to a nitrogen atom, oxygen atom, or fluorine atom. If there are no H-N, H-O, or H-F bonds, there's no hydrogen bonding. The big question, of course, is why are hydrogen bonds different than regular dipoledipole forces? After all, wouldn't a molecule with hydrogen bonding be polar? Of course it would! In the hydrogen-oxygen bond (or H-N or H-F bonds), the hydrogen is less electronegative than the other atom, causing the electrons in the bond to be be pulled away from it. However, in a hydrogen bond, you'll notice that the atoms that hydrogen is bonded to each have lone pairs. Because the oxygen, fluorine, and nitrogen atoms all have these lone pairs, the lone pairs tend to be attracted toward the partial-positively charged hydrogen atoms on nearby molecules. Check out the diagram below:

Hydrogen bonds are generally much stronger than dipole-dipole forces, because the interaction of the lone pairs with the hydrogen atoms on another molecule are very strong.

3) Van der Waals forces: Van der Waals forces are when nonpolar molecules stick together due to magnetic attractions. Now, I know what you're going to say: "Mr. Guch, only polar molecules behave like magnets. How can a nonpolar molecule have magnetic attractions? If a molecule is nonpolar, where does the magnetic charge come from? Furthermore, wouldn't this imply that nonpolar molecules are really polar? What gives?" Whoa there! Hold off on the attitude before I get angry.

OK... that's better. Here's a dumb little story to illustrate what happens.

These two molecules are both nonpolar. The electric charge in each molecule is represented by the color green. As you might imagine, the green color is even spread all over the molecule because none of the atoms grab electrons more than any of the others. These molecules are happy. Uh oh. There's a problem in happy molecule land. As you might remember, the electrons in atoms jump around pretty much at random within a specific area. Unfortunately for the atom on the left, the electrons have all jumped to the left of the molecule because of these random movements. This has caused the molecule on the left to get a little bit of polarity, as shown by the red plus and minus signs. Even more trouble is on the way. Because the molecule on the left became slightly polar, the negatively-charged electrons in the molecule on the right all moved over so they could be next to the partial positive charge on the molecule on the left. Because both molecules now have a little bit of polarity, they stick to each other. Fortunately, all is well in nonpolar molecule-land. Because these polarities depended on random electon motion, it goes away when the random electron motion makes the charges even out over the whole molecule. For this reason, this is referred to as "temporarily induced dipoles", because the dipoles only last for a very short time. Not only are these forces short-lived, but they're also very weak because they don't last for very long. Van der Waals forces are the weakest of the three intermolecular forces.

How can we tell the type of intermolecular force in a molecule?


Actually, it's not that hard to figure out whether a molecule will experience hydrogen bonding, dipole-dipole forces, or Van der Waals forces. All you have to do is the

following: 1) Draw the Lewis structure for the molecule. If you don't know how to draw Lewis structures, click HERE for the helpdesk section on how to do that. 2) Examine the Lewis structure and see if there's a hydrogen atom stuck directly to a fluorine, nitrogen, or oxygen atom. If it is, the molecule experiences hydrogen bonding. An important note, however, is that there may be some molecules that contain F, N, or O as well as hydrogen that aren't involved in hydrogen bonding - an example would be HCN, which has a Lewis structure that looks like HC:::N: Remember, hydrogen bonding only takes place when the H is stuck directly to F, N, or O. 3) If the molecule doesn't have a H-N, H-F, or H-O bond, you need to keep looking. The next question: Is the molecule polar? If it is, the molecule experiences dipole-dipole forces, and if it isn't, it experiences Van der Waals forces. Of course, the big question here is this: How can you tell if a molecule is polar? Hey, I was getting to that. Basically, you can tell if a molecule is polar by looking to see if the Lewis structure is completely symmetric. One way you can do this is to take a look at the central atom and see if the atoms around it are all arranged evenly. If they are, the molecule is nonpolar. If they aren't, then it's polar. Let's see some examples below:

Take a look at the molecule on the right. Maybe you're wondering why it's drawn like this instead of with both fluorines on opposite sides of the sulfur atom. After all, if we did that, wouldn't the molecule appear to be nonpolar? It would, and the molecule is actually polar, so that would give us a false impression. It's a general rule of thumb that if the central atom in a molecule (like the P, C, and S in the examples above) has a lone pair of electrons, the molecule is polar. I'm sure there are examples where this isn't true, but I can't think of any right now, so they can't be that common.

What's all this good for, anyway?


I'm glad you asked. As it turns out, intermolecular forces explain a lot of stuff about

how compounds behave. For example, if a compound experiences strong intermolecular forces such as hydrogen bonding, it will have a higher melting and boiling point. This is because hydrogen bonding helps to stick the molecules together, making it harder to pull them apart. Because energy is required to make this kind of change take place, more energy in the form of heat will be needed to make hydrogen-bonded molecules melt. Another thing that intermolecular forces can do is to explain the whole "like dissolves like" thing. You've probably heard this saying, and it means that molecules with similar polarities tend to dissolve one another. If you know how to find the intermolecular forces a molecule experiences, you can figure out whether one thing will dissolve another based on their polarities. Pretty handy! Finally, it's handy because your teacher will probably ask you to come up with intermolecular forces on a test. But you already knew that, I bet.

To do practice worksheets about polarity, click HERE. To go back to the Helpdesk, click HERE. To go back to the Cavalcade o' Chemistry main page, click HERE.

Questions? Comments? Angry rants? Email them to me at misterguch@chemfiesta.com

So, what's the deal with solutions?


Head back to the helpdesk Find more tutorials Try some practice worksheets

Covered in this discussion:


Explanation of what a solution is. Molarity: What it is and how to calculate it. Colligative properties: What they are and why we should care.

Explanation of what a solution is


Whenever you dissolve something in a liquid, you've made a solution. Another way of thinking about it is that if you have a liquid and it's not a pure substance, it's a solution. The liquid part is the "solvent" and the stuff you dissolved (usually a solid) is the "solute". An example: If you dissolve salt in water, the salt is the solute and

the water is the solvent. A handy rule of thumb: If somebody asks you to tell you what the solvent in a solution is and you have no idea, say "water". Water is by far the most common solvent for solutions that you're likely to run into. When we're talking about how much of the solute is dissolved in the liquid, we're talking about the concentration. There are four terms we can use to describe the concentration of a solution:

unsaturated: This means that if you were to add more solute to the liquid, it would keep dissolving. For example, if you take one teaspoon of salt and put it in a bucket of water, you've made an unsaturated solution. (In other words, if you added another teaspoon of salt, it would dissolve, too). saturated: This means that the liquid has dissolved all of the solute that is possible. If any of you have a little brother or sister who adds sugar to iced tea, you know what I'm talking about. If you add one teaspoon of sugar to iced tea, you've got an unsaturated solution. If you keep adding sugar to iced tea, you eventually get to the point where the rest of the sugar just sinks to the bottom. When this happens, it means that the solution is saturated, because no more sugar could dissolve. supersaturated: This means that MORE solute has dissolved than is possible. How, you might ask, does this happen? If you have a very hot saturated solution and cool it down, the solubility of the solute decreases as the solution cools. (In other words, hot solutions can dissolve more solute than cold ones). What usually happens in this situation is that the solute starts forming crystals at the bottom of the container. However, under weird circumstances where there are no little grains to start crystal formation, the crystals never form - as a result, the solution is MORE concentrated than possible. This doesn't happen much, so you'll never run into it in real life, most likely. molarity: This is another unit of concentration. Keep reading for more...

Molarity: What it is and how to calculate it


Molarity is one of those terms that people like to talk about a lot when describing solutions. Unlike "unsaturated", "saturated", and "supersaturated", molarity is a numerical way of saying exactly how much solute is dissolved in a solvent. As you can probably tell from the section before, the term unsaturated or supersaturated can be applied to a wide variety of solution concentrations. OK. Here's the definition of molarity: Molarity is equal to the moles of solute divided by the liters of solution. To put it in

the form of an equation:

OK. Quit freaking out. It's not that hard to figure out. Let's do an example: Example: What's the molarity of a solution that contains 5.5 moles of sodium chloride in 10.5 liters of solution? Answer: M = moles / liters, or (5.5 moles) / (10.5 liters) = 0.52 M. In this case, the unit is M. M stands for "molarity". A 0.52 M solution is referred to as being a "0.52 molar" solution. That's simple enough!

Sometimes these problems get a little bit harder. Instead of giving you moles, they give you grams. Instead of liters, they give you milliliters. Fortunately, you know how to do the necessary conversions. Here's a handy diagram to help you:

Let's do an example. Example: If I have 3.50 grams of sodium chloride in 1250 mL of a solution, what's the molarity? Solution: To find molarity, we need to convert grams to moles and milliliters to liters. To convert grams to moles, we first need to divide the number of grams by the molar mass of sodium chloride. (3.5/58.5) = 0.060 moles. To convert milliliters to liters, multiply by 0.001. (1250 x 0.001) = 1.25 liters. In the final step, divide the number of moles by the number of liters to get the molarity. Since 0.060 / 1.25 = 0.048, the molarity of the solution is 0.048 M. You are now a molarity expert!

Colligative properties
Not surprisingly, when you dissolve something in a liquid, it has different properties than the pure liquid. Any properties that change when the concentration of a solution changes are calledcolligative properties. People always list boiling point elevation, melting point depression, and osmotic pressure as the main colligative properties of liquids. We're going to do it a little differently - after all, I'm not trying to sell you guys a textbook, so I don't have to use the confusing words that school districts like to see. Think about Kool Aid. Let's make a weak Kool Aid solution by dissolving one grain of Kool Aid in a glass of water. Let's also make a strong Kool Aid solution by dissolving a cup of Kool Aid powder in a glass of water. The properties that are different between the two glasses of Kool Aid are "colligative properties". In our example, we would find that: Strong Kool Aid is darker than weak Kool Aid. As a result, we would say that "color" is a colligative property of liquids. This is the basis for a field called spectophotometry, where you can figure out how concentrated a solution is by looking at the color. Strong Kool Aid is a lot sweeter than weak Kool Aid. To make it a little broader, we can say that "taste" is a colligative property of liquids. True, but it's usually not all that handy in the lab. They say that sodium cyanide tastes like almonds, but I wouldn't test the concentration of a cyanide solution by taste! (Who figured that out, anyway?) Strong Kool Aid is thicker and goopier than weak Kool Aid. Texture is a colligative property. Strong Kool Aid is denser than weak Kool Aid. Density is a colligative property. Strong Kool Aid boils at a higher temperature than weak Kool Aid. Boiling point is a colligative property. Strong Kool Aid freezes at a lower temperature than weak Kool Aid (if you've ever tried to make homemade popsicles, you know this to be true). Melting / freezing point is a colligative property. Now that you know what colligative properties are, let's look at the ones that your textbook probably focuses on: Boiling point: As the molarity of a solution increases, the boiling point increases. This is because the solute lowers the vapor pressure of the solution and solutions don't boil until the vapor pressure of the solution is equal to atmospheric pressure. If this doesn't make any sense, just remember that strong solutions boil at higher temperatures than weak ones. Melting point: As the molarity of a solution increases, the melting point decreases. This is because the solute keeps the solvent molecules from forming a nice solid lattice. Think of this: Ocean water doesn't freeze very easily - this is

because it's got salt dissolved in it. Osmotic pressure: You should have heard about this one in your biology class somewhere along the road. Basically, if you've got a solution that's separated from a pure solvent by a semi-permeable membrane, the pure solvent likes to move across the membrane to decrease the concentration of the solution. The pressure that the solvent pushes across the membrane with is called osmotic pressure. Not surprisingly, the more concentrated the solution, the more the pure solvent likes to push across the membrane and the higher the osmotic pressure.

OK. That's it for solutions. I hope this has helped. If so, tell all your friends to come visit and learn about solutions. If not, then keep it to yourself because I don't need the bad publicity. If you have questions, email me at misterguch@chemfiesta.com. You can even ask me stupid questions if you'd like - I may laugh when I read them, but I'll still email you back. I'm that kind of guy. If you email me your homework questions, I'll know and make fun of you in my response. I'll still email you, though.

Fun With Colligative Properties:


Head back to the helpdesk Find more tutorials Try some practice worksheets

Not surprisingly, some of the properties of solutions change when you change the concentration. Concentrated solutions generally tend to be thicker, darker in color, and denser. Colligative property: Any property of a solution that changes when the concentration of the solution changes.

Examples of colligative properties and how they work:

Boiling point elevation: o As you increase the concentration of a solution, the boiling point of the solution increases.

o The reason for this is that adding solute decreases the vapor pressure of a solution because there are fewer molecules at the surface of the solution that can vaporize. Because liquids boil when their vapor pressure is equal to the ambient atmospheric pressure, you need to heat them more to make them boil. o The relationship between concentration and boiling point is described by the equation:

Tb = Kbm
Where: Tb is the change in boiling point Kb is the ebullioscopic constant (0.520 C/m for water) m is the effective molality of the solute o Effective molality refers to the molality of particles in the solution. For covalent compounds, the effective molality is the same as the regular molality because the molecules dont break into smaller pieces when you put them in water. For example, if you have a 0.75 m solution of CS2, the effective molality is 0.75 m.

For ionic compounds, the effective molality is equal to the molality times the number of ions in the compound. For example, if you have a 0.75 m solution of NaCl, the effective molality is 1.50 m (0.75 x 2) More examples, if needed.

o Sample problems: What is the boiling point of a 1.75 m solution of H2CO? 100.910 C What is the boiling point of a 1.75 m solution of NaCl? 101.820 C What is the boiling point of a 1.75 m solution of Ca3(PO4)2? 104.550 C

Melting point depression: o Solutions melt at lower temperatures than pure liquids because the solute molecules disturb the intermolecular forces between the solvent molecules. As a result, less energy is needed to break up the solid. o The relationship between concentration and melting point is determined by the equation:

Tf = Kfm
Where: Tf is the change in freezing point Kf is the cryoscopic constant (1.860 C/m for water) m is the effective molality of the solute o The rules for determining effective molality are the same as for boiling point. The only thing different is the constant used. Example: What is the melting point of a 1.75 m solution of NaCl? -6.510 C

Vapor pressure decrease o The more concentrated a solution, the lower the vapor pressure (for the reasons we described above with the BP elevation):

Osmotic pressure increase o Osmotic pressure is the force with which a pure solvent moves across a semi-permeable barrier into a container that holds a solution (explain what a semi-permeable barrier is). o This occurs because solvents tend to diffuse across barriers in ways that will decrease the difference in concentration. The bigger the difference in concentration (because of high solute concentration), the stronger the force of osmosis.

Conductivity of electricity o Electrolytes: Solutions that are able to conduct electricity. These are solutions in which ionic compounds are the solute recall that ionic compounds are good at conducting electricity when dissolved in water.

o The more concentrated the solution of an ionic compound, the better it is able to conduct electricity. The larger the number of mobile ions present, the larger the number of electrons that can be moved from one place to another.

2009 Cavalcade Publishing and Ian Guch, may not be reproduced without permission. For more information, email Ian Guch at misterguch@chemfiesta.com

Acids 'n' Bases


Head back to the helpdesk Find more tutorials Try some practice worksheets

On this tutorial page you can find the following:


Definitions of acids and bases Properties of acids and bases Discussion of the pH scale and pH calculations Titrations Miscellaneous other stuff

What are acids and bases? Definitions and Properties


You hear about acids all the time. Acid reflux disease causes some people to have

to take acid reducing medication. The fact of the matter is that you hear the word "acid" all the time. Most of us, however, don't have any idea what an acid is. The Bronsted-Lowry definition of acids is that acids are compounds that give off H+ ions when they react with another compound. Likewise, this definition says that bases are compounds that accept H+ ions from other compounds. The Arrhenius definition of acids says that they're compounds that give off H+ ions in water and that bases are compounds that give off OH- ions in water. These definitions are the same. Basically, if you've got something that can give off H+ in water, it's an acid. As a result, all acids you'll be seeing in class have hydrogen atoms on them that are ready to go jumping off in water. Most common acids have the letter H in the beginning of the formula, with the exception of acetic acid (it's at the end, for reasons we won't go into here). Bases, on the other hand, are compounds that give off OH- in water. (The two definitions of a base are for our purposes identical, as OH- combine with H+ to form water -- the Arrhenius and Bronsted-Lowry definitions are for most purposes identical). When you see the formula of a base, it's got "OH" in it. The one exception to this is ammonia, NH3. (NH3 combines with water to form NH4OH, which is really the thing that's basic in ammonia. So our definition is sort of true). You can also define acids and bases as being "strong" or "weak". Strong compounds are compounds that completely break up in water. In other words, if we're talking about a strong acid, all of the H+ ions break away from the molecule in water. For strong bases, all of the OH- ions break away from the molecule in water. There is a difference between a "strong" acid and a "reactive" one. Strong acids are all reactive, but some "weak" acids can also be extremely reactive. A good example of a weak, reactive acid is hydrofluoric acid, HF. I had a friend of mine who had a tube full of HF explode in his face - even though it's a weak acid, he still spent a long time recovering and suffered permanent scarring. Ask your teacher sometime which acid they'd rather put their hand into, HCl (a strong acid) or HF (a weak acid). If your teacher knows anything at all about acid chemistry, they'll reply HCl. Here are a couple of charts which show the most common acids and bases. Some are strong and some are weak, as indicated.

Acids
Formula HCl HBr Name hydrochloric acid hydrobromic Strong? yes yes Formula NaOH LiOH KOH

Bases
Name sodium hydroxide lithium hydroxide potassium Strong? yes yes yes

acid HI HF HNO3 H2SO4 H3PO4 CH3COOH hydroiodic acid hydrofluoric acid nitric acid sulfuric acid phosphoric acid acetic acid yes no yes yes no no Mg(OH)2 Ca(OH)2 NH3 (NH4OH)

hydroxide magnesium hydroxide calcium hydroxide ammonia (ammonium hydroxide) no no no

Properties of acids and bases


Properties of acids include the following: React with most metals to form hydrogen gas Taste sour (like lemons) Frequently feel "sticky" Usually gases or liquids Properties of bases include: Feel "slippery". (FYI: The slippery feeling is caused because your skin dissolves a little when you touch them.) Taste bitter (like baking soda) React with oils and greases (that's why they're used as drain and window cleaners) Frequently solids (though ammonia is a gas that's usually dissolved in water)

The pH Scale
Everybody has heard of pH. You've seen it in middle school, you've heard people talk about it in shampoo commercials, and you can even buy deodorant that's "pH balanced", whatever that means. Unfortunately, most people don't know what pH is. pH is a measurement of the H+ concentration in a liquid. If there's a high H+ concentration, the pH indicates that you've got a very acidic solution. If the solution is neutral, there's only a small H+ concentration, and the pH reflects that. If the solution is basic, there's almost no H+ concentration, and you can tell that by the pH number. pH is nothing more than a way of telling how concentrated an H+ solution is. Here's something people always have problems with: If I have a neutral or basic

solution, how come I can measure the pH? After all, you can only measure pH if there's H+ ions present in a solution, and shouldn't the H+ concentration be 0 for neutral and basic solutions? Good question. It turns out that water has the funny property that it tends to spontaneously break up into H+ and OH- ions no matter how much acid or base you've added. As a result, even very basic solutions have a little bit of H+ floating around in it. You'd be right in guessing that a very basic solution doesn't have much H+ because the OH- in the base reacts with it to form water. In any case, not much of the water breaks up like this: In neutral solutions the concentration of H+ is only 10-7 molar. (Even though there's a little bit of acid present, the solution is still neutral because there's also a little bit of OH- present - in neutral solutions, this is also 107 molar). Now that I've described what the pH scale is, let's take a look at it:

As you can see, pH values between 0 and 7 are acidic and pH values between 7 and 14 are basic. pH values of exactly seven are called "neutral" solutions - if the pH is 6.99 it's an acidic solution and if it's 7.01 it's basic. However, people usually refer to solutions with a pH between 6 and 8 as being "neutral" because they're mostly neutral. Don't put this on a test, though, because it will be marked WRONG as it's technically wrong.

pH Calculations
OK. We're getting to the part with the calculations here. Now that we know what the pH scale is, let's learn how to compute the pH of a solution. The equation you need for these calculations is simple:

pH = -log[H+]
In this equation, [H+] refers to the molarity of the acid in the solution you're looking at. As a result, you'll need to be given (or calculate) the concentration of acid present before you can do this problem. Once you know this, it's just a matter of

plugging this equation into your calculator. I'll show you how to do this in an example. Example: What's the pH of a 0.05 M solution of HCl? Solution: To solve this, you only need to realize that in this case [H+] = 0.05. After that, it's just a matter of plugging this thing into your calculator. Let's learn how to do that.

How to solve this equation if you have a big fancy graphing calculator: 1) There's a button that looks like this at the bottom of your calculator: (-) [Note: On some calculators it may look like "+/-". Don't just hit the minus sign at the right of your calculator, because that'll give you an error. 2) Hit the "log" button on your calculator. You should now see something that looks like this on your screen: -log( 3) Type the concentration into the calculator. In our example, this is 0.05. You should now see this: -log(0.05 4) Hit the ")" button on your calculator and hit the "enter" button. Your calculator should now display the correct answer, 1.30. How to solve this equation if your parents won't buy you a big fancy graphing calculator: 1) Type the concentration into your calculator. In this case, it's 0.05. 2) Hit the "log" button. 3) Hit the button on your calculator which looks either like "(-)" or "+/-". 4) There's your answer! As you can tell, you've got it easier than those people who have to type all that junk into their fancy graphing calculators. Make sure to tease them while they're fiddling with all those fancy buttons.

If you don't have a scientific calculator, you're kind of out of luck, unless you have a log table (which nobody ever does, because if you need a log table you usually go buy the $10 calculator rather than the $30 log table, which is a lot less useful). As your math teacher if you can borrow a scientific calculator, borrow one from a friend, or keep bugging your parents until they give you ten bucks to buy a scientific calculator. For the record, don't tell your parents that I told you to get a graphing calculator - in my opinion, graphing calculators are useless for chemistry classes because you never actually need to graph anything in calculator format that you wouldn't also write down on a piece of paper.

Titrations
You've heard this word before. Come on, admit it. Your teacher talked about this in class and you didn't have any idea what he/she was talking about so your decided to pretend like it didn't exist. Well, I'm here to tell you it does. Might as well keep reading. Titrations are not all that hard to understand. In fact, the word "titration" comes from the Greek titros which means "to figure out the molarity of an acid or base solution" and the Latin ationswhich means "by neutralizing it with a solution whose concentration you already know". Those ancient people really had a way with words. Here's the idea. Let's say that you had really bad eyes and wanted to see how many toothpicks you had in a pile. In fact, your eyes are so bad that you can't even see the toothpicks to pick them up, much less count them accurately. This poses a problem. Your friend has an idea. You've got a bunch of little sandwiches lying around the house from the dinner party your parents hosted last night. If your friend sticks one toothpick into each sandwich, you could figure out how many toothpicks you had because all you'd need to do is count the number of sandwiches. You wouldn't be measuring the number of toothpicks directly by counting them, you'd be measuring them secondhand by how they interacted with something else. That's what a titration is. Let's say you have an acidic solution and wanted to figure out the molarity. Well, you can't do that directly, because you can't count acid molecules. They're too small. You can, however, make a basic solution with a concentration that you already know. If you keep adding base to the acid, eventually all of the acid molecules will be neutralized and the solution will turn from an acid to a base. If you know how many base molecules you added to the solution before the solution gets neutralized (and you will, because you'll add the solution drop-by-drop), you can figure out how much acid was in the solution in the first place. Of course, this leads to an interesting problem: How can you tell when the solution gets neutralized? The answer: Indicators! Indicators are chemical compounds that turn different colors when they're in solutions with different pH's. The indicators you'd most likely work with turn color when the solution becomes neutralized. Litmus, for example, is red in acid solutions and blue in basic solutions. Phenolphthalein (pronounced fee-no-thay-leen) is clear in acid solutions and pink in basic solutions. OK. Now that you have the basic idea behind titrations and know what indicators are, let's figure out how to solve some problems. The basic equation you need is this:

M1V1 = M2V2

M1 stands for the molarity of the acid V1 stands for the volume of the acid you use M2 stands for the molarity of the base V2 stands for the volume of the base you use

Let's do an example that might make make this more clear. Example: If it takes 55 mL of 0.1 M NaOH solution to neutralize 450 mL of a HCl solution of unknown concentration, what's the molarity of the acid? Solution: Before you can do anything, you need to translate this into something that makes sense. Let's go through it slowly.

M1, in our equation, stands for the molarity of the acid. Since that's what we're trying to find, we'll call that X. V1 stands for the volume of the acid we use. Since HCl is an acid, the volume of acid is 450 mL M2 stands for the molarity of the base. Since NaOH is a base, the molarity was stated in the problem to be 0.1 M V2 stands for the volume of the base. The problem says that we used 55 mL of base, so that's M2. Now, all we need to do is plug it into the equation: (X)(450 mL) = (0.1 M)(55 mL) X = 0.12 M And that's your answer!

Miscellaneous other stuff


In this section, I'm just going to talk about a bunch of other miscellaneous stuff that's a little more advanced. I'm not going to pretend that these things are especially complete or complex - they're mainly just here to fill in some of the main ideas I may have missed. If I don't talk about something in detail, it's either because I couldn't figure out a good way of explaining it over the computer, it's obscure enough that I don't think that all that many people would really need to know about it, or I just forgot about it when I was writing this section up. Whatever the reason, if it ain't here you should ask your teacher for help. Conjugate acids and bases You already know what an acid and base are. To put it very simply, acids have H+ in them and bases have OH- in them. One thing we didn't talk about above is that

every acid has a conjugate base and every base has a conjugate acid. You may be asking yourself, "What the heck does that mean?" Well, settle down Florence, we're getting to it. The conjugate base of an acid is whatever is formed when the acid loses its H+. An example: If HCl loses the H+, you end up with Cl-. As a result, Cl- is the conjugate base of hydrochloric acid. It's as simple as that. As a general rule of thumb, the conjugate bases of strong acids are weak. For example, Cl- is the conjugate base of hydrochloric acid. If you wanted to, you could go swimming in a large body of water containing lots of Cl- and never get hurt. As a matter of fact, the ocean is full of Cl- (it comes from NaCl). So are you. Buffers Buffers are solutions that don't change pH very much when you add acid or base solutions to it. For example, if you were to add a little bit of HCl to a glass of water, the pH might change from 7 to 3. If you had the same amount of buffer solution, the pH might change from 7 to 6.8. Neat stuff. Buffers are formed when you have a weak acid and its conjugate base present in the same place. If you wanted to make an acidic buffer, you'd place some acetic acid into a container with some sodium acetate. Voila! You have a buffer. If you want a basic buffer, just put a weak base into a container with it's conjugate acid. Same deal. Your blood is a buffered solution. If it wasn't, your pH would be go way down every time you had a soda and way up whenever you took some Tums. It's not a very handy survival technique to die every time you have a soda.

As my favorite band says at the end of an album, "That's all the singing." If you've got any questions or comments, email them to me at misterguch@chemfiesta.com and I'll do my best to get back to you.

The Gas Laws


Head back to the helpdesk Find more tutorials Try some practice worksheets

Lots of students have trouble when doing gas law calculations. Typically, these problems aren't so much about doing the math once you have all of

the right variables. Instead, the problem involves figuring out which equation to use for any particular problem. In this section of the Helpdesk, we'll discuss how to figure out which gas laws need to be used, as well as how to use them and common pitfalls that people encounter doing gas laws calculations.

Getting started: some handy terms you'll need to know Before doing gas law calculations, you've got to figure out what all of the appropriate terms, symbols, and variables are. For those of you having trouble figuring it out, here are some of the more common ones: ideal gas: An imaginary model of a gas that has a few very important properties. These properties are that the particles of the gas are assumed to be infinitely small, the particles move randomly in straight lines until they bash into something (another gas molecule or the side of whatever container they're in), the gas particles don't interact with each other (they don't attract or repel one another like real molecules do) and the energy of the particles is directly proportional to the temperature in Kelvins (in other words, the higher the temperature, the more energy the particles have). We make these assumptions because a) They make the equations a whole lot simpler than they would be otherwise, and b) Because these assumptions dont' cause too much deviation from the ways that actual gases behave. kelvins: A temperature scale in which the degrees are the same size as degrees Celsius but where "0" is defined as "absolute zero", the temperature at which molecules are at their lowest energy. To convert from degrees Celsius to Kelvins, add 273. By the way, we don't say "degrees Kelvin", we just say "Kelvins". Go figure. pressure: A measure of the amount of force that a gas exerts on whatever container you've put it into. Imagine a shaving cream can. When the pressure is very high in there, the gas in the can pushes very hard on the walls of the can, which is why the cans are made much stronger than soda cans. Units of pressure include atmospheres (1 atm is the average atmospheric pressure at sea level), Torr (which are equal to 1/760 of an atmosphere), millimeters of mercury (1 mm Hg is the same as 1 Torr, or 1/760 atm), and kilopascals (there are 101.325 kPa in 1 atm). standard temperature and pressure : A set of conditions defined as 273 K and 1 atm. temperature: A measure of how much energy the particles in a gas have. Units of temperature that you'll run into include degrees Celsius (which you shouldn't use when doing gas law calculations for reasons we'll talk about later) and Kelvins (which is equal to 273 plus the degrees Celsius). volume: The amount of space that some object occupies. The unit of volume can

be cubic centimeters (abbreviated "cc" or "cm2"), milliliters (abbreviated "mL" - 1 mL is the same as 1 cubic centimeter), liters (abbreviated as "L" and equal to 1000 mL), or cubic meters (abbreviated "m2" - there are one million cubic centimeters in a cubic meter). Avogadro's Law Amadeo Avogadro was a man with a dream. He knew that if he studied and practiced and worked really hard, he could be the world's most famous tuba player, renowned throughout the modern world for all time. Tragically, his dream died during a freak tuba accident in the early 1800's - though he was able to regain the use of his tuba hand, it never recovered enough for him to gain superstar status*. Heartbroken, he turned to science, coming up with a law that we use today, called, straightforwardly enough, Avogadro's Law. Here's what Avogadro said: If you have the same volumes of two gases, they'll have the same number of molecules. For example, one liter of carbon dioxide will have just as many molecules in it as a liter of nitrogen. He figured this out through a series of experiments using the very best equipment of the day. Of course, this equipment wasn't all that great and the error masked the fact that this law isn't, in fact, true. However, it's mostly true, which allows us to assume that all gases (roughly) behave the same under the sam conditions of volume, temperature, and pressure. Without this law, gas law calculations would be very, very inconvenient.
* The tuba story is a total lie. I made it up because I got bored of only talking about gas laws.

Boyle's Law Robert Boyle was another man with a dream. He wanted to be the first man to eat 100 hard boiled eggs in a 24-hour period. Unfortunately, some of the other chemists got jealous - let's just say that considerable ugliness ensued and Boyle's dream was permanently derailed. However, Boyle was a man of many talents, and was able to come back from his humiliating egg fiasco* to come up with a gas law of his own. Here's what Boyle did: He put a gas into a container in which he could change the volume and measure the pressure. When he multiplied the volume of the gas times it's pressure, he found it was equal to some arbitrary number (let's call it k, because he did). If he changed the pressure of the gas, he found that the volume also changed, which isn't really surprising (if you push on something, it gets smaller). What is surprising is that if you multiply the new pressure by the new volume, the answer is the same arbitrary number that you had in the first place (k!). From this, we can make the following statement: P1V1 = P 2V2 In this equation, P1 is the initial pressure of the gas and V1 is the initial volume of the

gas. P2 is the final pressure of the gas and V2 is the final volume of the gas. This way, if you know the initial pressure and volume of a gas and know what the final pressure will be, you can predict what the volume will be after you put the pressure on it. Let's see an example.

Question: If we have 4 L of methane gas at a pressure of 1.0 atm, what will be the pressure of the gas if we squish it down so it has a volume of 2.5 L? Answer: Let's plug the numbers we've been given into the problem. P1 is 1.0 atm and V1 is 4 L. After we squish the gas, the volume (V 2) is 2.5 L. When we put all of these numbers into the equation, we get: (1.0 atm)(4 L) = (x atm)(2.5 L) x = 1.6 atm

* I also made up the egg incident.

Charles' Law Jaques Charles was a disturbing and scary guy. Though he came up with a really handy law for determining what the relationships between the volume and temperature of a gas are, his private life was far more bizarre. Some say that if you go by the old Charles mansion at the edge of town, you can still hear the moaning and wailing of his ghost, forever roaming the night.* Anyhow, what Charles determined through his studies was that when you change the temperature of a gas, the volume changes. Not surprising - you probably know already that if you heat something, it tends to get bigger. What he found, though, was that if you divide the volume by the temperature of a gas at one temperature, you get a constant. Just like Boyle found, if you change the volume or temperature of this gas, you get the same constant. From this, Charles came up with this statement: V1/T1 = V 2/T2 Where the subscript "1" indicates the initial volume and temperature and the subscript "2" indicated the volume and temperature after the change. Temperature, incidentially, needs to be given in Kelvins and not in Celsius - this is because if you have a temperature below zero degrees Celsius, the calculation works out so the volume of the gas is negative, and you can't have a negative volume. Let's see an example of this equation in action:

Question: If we have 2 L of methane gas at a temperature of 40 degrees Celsius, what will the volume be if we heat the gas to 80 degrees Celsius? Answer: The first thing we have to do is convert the temperatures to Kelvins (by adding 273), because Celsius can't be used in this equation. To do this, we get that the initial temperature is 40 + 273 = 313 K and the final temperature is 80 + 273 = 353 K. We're now ready to start sticking these numbers into the equation: 2 L / 313 K = x L / 353 K x = 2.26 L
*This story isn't true.

Gay-Lussac's Law There was a third guy whose gas law is a little less famous than the others. Some think it's because of his funny name, while others think it has something to do with having a bad public relations firm working for him. Whatever the reason, his name is Gay-Lussac, and his law related pressure to temperature:

P1/T1 = P2/T2
This gas law explains how if you increase the temperature of a container with fixed volume, the pressure inside the container will increase. This explains why you shouldn't leave cans of spray paint in your trunk - the pressure might get so high that the propellant will blow the can up. The Combined Gas Law Imagine a world in which you didn't need to memorize the three laws above. Instead, there was one big law that covered both of them. Hey, that's the world you live in now, and the law you need to know is the combined gas law: (P1V1) / T 1 = (P2V2) / T2 In this equation, all of the terms are exactly the same as in the preceding equations. The way you can use this equation is that whenever you're changing the conditions of pressure, volume, and/or temperature for a gas, you just plug the numbers into this equation. However, let's imagine that the temperature of the gas didn't change while you were making your change. Since the first temperature term and the second are the same, they cancel out. As a result, if one of these variables isn't mentioned in the problem, just ignore it entirely. Let's see an example:

Question: If we have two liters of a gas at a temperature of 420 K and decrease the temperature to 350 K, what will the new volume of the gas be?

Answer: To solve this problem, use the combined gas law to find the answer. Since pressure was never mentioned in this problem, just ignore it. As a result, the equation will be: V1/T1 = V 2/T2 Which is the same thing as Charles's law. To solve, the initial volume is 2 L, the initial temperature is 420 K, and the final temperature is 350 K. The final volume, after solving the equation, should be 1.67 L. The Ideal Gas Law What happens if you don't change the conditions of a gas, but just want to find out what a gas is like when it's sitting in a container, not doing much? Well, the equations above won't help you much, because they're equations which depend on making a change and comparing the conditions before the change and after the change to make determinations about what the gas is like. The ideal gas law is an equation of state, which means that you can use the basic properties of the gas to find out more about it without having to change it in any way. Because it's an equation of state, it allows us to not only find out what the pressure, volume, and temperature are, but also to find out how much gas is present in the first place. Here it is: PV = nRT Where P is the pressure of the gas (either in atmospheres or kilopascals), V is the volume (in liters), n is the number of moles, R is the ideal gas constant, and T is the temperature (in Kelvins). There are two common values for the ideal gas constant. One of them is 0.08206 L x atm / mol x K , and the other is 8.314 L x kPa / mol x K. The question is, which one do you use? The value of R used depends on the pressure given to you in the problem. If the pressure is given to you in atmospheres, use the 0.08206 value because the unit at the end of it contains "atmospheres". If the pressure is given to you in kilopascals, use the second value because the unit at the end contains "kPa". Another good thing about this law: It allows us to figure out how many grams and moles of the gas are present in a sample. After all, "moles" is the "n" term in the equation, and we already know how to convert grams to moles (if you've forgotten how, click here ). Let's see an example:

Question: If I have 4 liters of a gas at a pressure of 3.4 atmospheres and a temperature of 300 K, how many moles of gas are present? Answer: The first thing you need to do is figure out what value of the ideal gas constant should be used. Since pressure is given to you in "atmospheres", use the first one, 0.8206 L x atm / mol x K. After plugging in the given terms for pressure, volume, and temperature, you end up with: (3.4 atm)(4 L) = n (0.08206 L x atm / mol x K)(300 K) n = 0.55 moles And it's as easy as that! For a practice worksheet with the ideal gas law, click here .

There's a whole bunch of other gas laws, but since I've found that students usually have troubles figuring out these, I've decided to stick to these. Well, that, and because my fingers got tired of typing. You know how it goes. Questions? Comments? Love letters? Angry rants? Email them to me at misterguch@chemfiesta.com

Energy Diagrams
Head back to the helpdesk Find more tutorials Try some practice worksheets

If you're looking for a Romanian translation of this page, you can find it here. Here's what a sample energy diagram looks like (this is for an exothermic reaction):

Important Terms
.

reagents: What you start with in a reaction. products: What is formed in the reaction. activation energy: The amount of energy needed to make the reaction take place. Usually abbreviated Ea. transition state: The midway point in the chemical reaction. Usually abbreviated Ts. heat of reaction: The amount of energy change between the reagents and products. Usually abbreviated H.

Now, let's talk about what all that stuff means!

What happens during a chemical reaction


Chemical reactions require a bunch of stuff to take place before they can occur. Let's talk about all of the steps here, and what they mean in terms of the energy diagram above. What needs to be done to start a chemical reaction: Before a chemical reaction can take place, we need to do a few things. First of all, we need to assemble the ingredients that we want to make form our products. These ingredients are calledreagents (sometimes referred to as "reactants"). Obviously, if we don't have ingredients, we won't be able to make anything! Sometimes just putting ingredients together is enough to start a reaction. However, in a lot of cases, we need to add a little bit of energy to get the reaction going. This

isn't really surprising - after all, if you bake a cake, it's not enough to make the batter and just wait around for the reaction to occur. You need to stick the batter in the oven so enough energy goes into it to make it into a cake. Well, the same thing happens with many chemical reactions - if you want the reagents to make your desired products, you'll need to add some energy to get the process moving. This energy is called the activation energy, because the chemical process won't activate until the energy is added to it. Some reactions have higher activation energies than others, but all have an activation energy of one type or another. For example, when you put gasoline in contact with oxygen, you know that you shouldn't let any sparks near it. This is because the activation energy for the combustion of gasoline is very low, and even the smallest amount of energy will get it started. On the other hand, if you want to make calcium carbonate turn into calcium oxide and oxygen, you need to heat it to about a thousand degrees Celsius. Because a lot of energy is required to make this reaction start, we'd say it has a very high activation energy. Reactions that proceed quickly generally have lower activation energies than slow reactions.

We're halfway there! The transition state: When we do a chemical reaction, the molecules that react come into contact with each other in such a way that they react. When the molecules are stuck in just the right way, they start to make the desired chemical change. The point at which the molecules have done exactly half of the change is referred to as the transition state. A good analogy to the transition state takes place whenever you go to the store to buy a candy bar. What happens is that you give the cashier the candy bar and they scan it on that little laser thingee. At some point during the transaction, you hand the cashier the money for the candy bar and he gives you the candy bar. That very instant in which you're giving the money to the cashier with one hand while grabbing the candy bar with the other is the transition state. At that instant, the cashier has halfway grabbed your money and you've halfway grabbed your candy bar. From that point on, the transaction practically completes itself. In chemical reactions, once the reagents have gotten past the transition state, they move smoothly to become products. Reactions with high activation energies occur more slowly than reactions with low activation energies. This happens because to make these reactions occur requires that the reagents have a lot of energy before they're able to reach the transition state. To go back to the candy bar analogy, imagine that you're trying to buy a candy bar from your worst enemy, and because he hates you so much, he doesn't want to sell it to you. If you're able to buy the candy bar at all, it'll be a slow process, because you'll have to expend a lot of energy trying to convince your enemy to sell you the candy bar in the first place. Eventually, if you spend enough time and energy, you'll be able to buy your candy bar - however, this will still take a long time.

Now, imagine that you don't want to spend a lot of time getting your enemy to sell you a candy bar. Is there a way that you can get the candy bar more easily? Sure! Just get a friend your enemy doesn't hate to buy it for you. Because your friend can buy the candy bar without all of the hassle you have to go through, the transaction will take place more quickly. However, the transition state for this reaction will be different - instead of you handing the money to the cashier, your friend will be handing money to the cashier. This kind of situation is exactly the same thing that occurs when a catalyst is used to speed up a chemical reaction. For one reason or another, some chemical reactions simply don't take place quickly. As a result, it's handy to speed the reaction using a catalyst whenever possible. What the catalyst does is to help the reagents combine in such a way to become products, frequently by sticking them together in just the right way to react. Though the same products will be formed, the transition state is different because the catalyst is involved in the transaction. When the reaction is over, the catalyst is unchanged. It didn't react, it just helped along the chemical reaction. Take a look at this picture for a better idea of what a catalyst does:

The exciting conclusion: Formation of the products After you've gone past the transition state in a reaction, it pretty much goes right on down to products without too much trouble. After all, for the reaction to go backwards, you'd have to climb from the products back to the transition state, which would require a lot of energy again. Let's face it, would you want to return the candy bar to your enemy after having gone to all the trouble of convincing him to sell it to you? Of course not! If the products have lower energy than the reagents, the reaction is said to be exothermic. Exothermic reactions give off heat, causing the things that were formed to have less energy than they did before they reacted. It may seem weird to

say that something that gives off heat would have less energy than something that doesn't, but let's consider yet another analogy. Let's say that for some reason or another, you've decided to go lie on a block of ice in your swimsuit. From the perspective of the block of ice, you're awfully warm. However, from your perspective, you're losing energy like crazy to the block of ice. When the ice has melted all the way, you'll have a lot less energy than you did before. That's how something that feels hot loses energy. If the products have greater energy than the reagents, the reaction is endothermic. Endothermic reactions absorb heat, making them feel cold. Again, let's say that you're lying on this block of ice. From your perspective, the block of ice feels very cold. Why is this? It's because the ice is absorbing all of your energy from you. The water actually has more energy than it did before you sat on it in ice form, because it absorbed all that energy from your body. Generally, endothermic reactions take place more slowly than exothermic reactions, because their activation energies are higher. Heat of reaction, or H, is just a measure of how exothermic or endothermic a reaction is. A reaction that's very exothermic would have a very negative H value, while a reaction that's only slightly exothermic would have a slightly negative H value. Heats of reaction aren't affected at all by the addition of a catalyst to the reaction, because no matter how it happens, the result of the reaction is the same. After all, if your friend buys a candy bar for you, in the end, you still have the same candy bar as you would have had if you'd argued with your enemy to get it.

Here's something for you to do to see if you're a pro: Sketch the energy diagram for an endothermic reaction, then compare it to the energy diagram that you can find by clicking HERE. If they're the same, then you know what you're doing!

Questions? Comments? Email them to me at misterguch@chemfiesta.com Want to see something stupid and amusing? Click here.

How can you identify an unknown substance?


Head back to the helpdesk Find more tutorials Try some practice worksheets
Table of Contents for this page:
1. When might you come into contact with unknown chemicals in the real world?

2. Simple tests you can do 3. Chromatographic methods What is chromatography? Thin layer chromatography and paper chromatography Gas-liquid chromatography (GC, or GLC) Other exotic chromatographies

4. Spectroscopic methods What is spectroscopy? Infrared spectroscopy UV-visible spectroscopy (UV-vis) Nuclear Magnetic Resonance (NMR) spectroscopy

5. X-Ray crystallography (a.k.a. X-ray diffraction, or XRD) 6. Mass spectrometry

1. Where might you come into contact with unknown chemicals in the real world? It turns out that the ability to identify unknown chemicals is a pretty handy ability to have. For that reason, people have spent unbelievably huge amounts of money learning how to do it. Let's say that you're working at a doctor's office, and some guy comes into the office gasping for air, claiming that he's been poisoned. What are you going to do? Well, depending on what he's been poisoned with, you're going to do different treatments. To maximize the effectiveness of the treatment, you need to figure out what he swallowed, or breathed, or whatever. Let's say that you're dropping a bunch of garbage off at the city dump, and you see a leaking barrel with a sign that says:

Danger! These chemicals are some real bad stuff! What are you going to do to get those chemicals cleaned up? It depends on what chemical they are. Let's say that the water you're drinking turns green and starts smelling funny? How are you going to clean up your water supply? Before you can do that, you've got to

figure out what's contaminating the water. Well, you get the idea. Let's talk about some ways that scientists figure out what unknown chemicals are. 2. Simple tests you can do OK. I've convinced you that identifying unknown chemicals is important. Let's talk about some methods of identifying unknowns. The following methods are pretty simple, and most can be done in even the crummiest chemistry lab. However, before we go too far, I want to interject a warning: Don't screw around with chemicals at home! ONLY screw around with them in chemistry class! Now that I said that, let's move to the simple methods of identifying unknowns:

Density: Density is equal to mass divided by volume. If you're trying to choose between two chemicals, you might want to take the density of the unknown by weighing it on a balance and taking the volume in a gradated cylinder. The chemical with the matching density is your unknown. Color: It's strange to think of something like color being an indicator of what an unknown is, but this is sometimes used by people who are very familliar with the unknown. However, as you might imagine, many chemicals look the same color, so this is only handy in some very limited situations. Smell: Smell can be a handy way of identifying a chemical if you've got a good nose. However, since many chemicals are extremely toxic, this is probably not a test you should use. Melting and boiling points: Most chemicals have melting and boiling points that are very different from others. If you can accurately find the melting and/or boiling temperature of your unknown chemical, you can probably match that value to a table of melting and boiling points to identify your unknown. The big problem with this method: If your chemical is contaminated, this will affect the melting and boiling point. You might also find that it's hard to get really good data in a laboratory setting. Let's look at some more conclusive and elaborate tests

3. Chromatographic methods What is chromatography? Chromatography is when you take a chemical and dissolve it in a gas or liquid (referred to as the carrier gas or liquid). You then run the carrier with the dissolved unknown over a surface in which somebody has stuck particles of a solid (the "stationary phase"). The idea here is that when you run the liquid over the particles,

some of the dissolved stuff will stick to the particles sbetter than the rest. By doing this, you can separate the compounds in a mixture, and hopefully figure out what they are. So, what makes the dissolved chemical stick to the particles? Well, one popular thing people take advantage of is similarities in polarity. (Polarity is just a measurement of how unevenly the electrons in a compound are arranged - compounds that have a very uneven distribution of electrons are referred to as being "polar"). Since compounds that are polar like to attract other polar compounds, the chemical that is more polar in a mixture will take longer to pass over the particles, because it will like to stick to them better. Still confused? Here's an analogy: Let's say that me and my great-aunt Gertrude go to the mall. I don't like the mall much, so I don't go into many stores. Gertrude, on the other hand, does nothing but shop, so she spends a ridiculous amount of time in the stores. Even though we started at the same point in the mall, we won't get to the other side of the mall at the same time, because she's stuck in the stores shopping, while I spend very little time in the stores. Chemicals do the same thing: If two chemicals start at the beginning of the solid particles, the one with less polarity (think of it as the one who likes to shop less) will pass through it more quickly than the one with higher polarity (the one that likes to shop more). That's how chromatography is done. Here's a sampling of some forms of chromatography which are commonly used:

Thin layer chromatography (TLC) TLC is a form of chromatography where the solid particles are just tiny pieces of silica gel (a polar compound) stuck to a piece of glass. What you do is take the mixture you want to separate and place it on the bottom of the glass plate, which only has one end in the nonpolar liquid (mobile phase). When the liquid runs up the plate, the nonpolar compounds will move farther up the plate than polar ones. TLC is most commonly used as a test to see how many chemicals are in a mixture. Usually, if you want to separate large amounts of a chemical, you need to do another form of chromatography. By the way, many classes use paper chromatography. This is exactly like TLC, except the solid particles are just paper fibers. Aside from that, the technique is the same.

Column Chromatography Column chromatography is the same as TLC, except that you fill a glass column with

silica gel particles and place the mixture on top. You then push a nonpolar liquid through the column to separate the compounds. Again, nonpolar compounds pass through the column faster than polar ones. Usually, the liquid is collected at the bottom of the column after it has passed through. With some skill and luck, you can separate compounds that are very close to each other in polarity.

Gas Chromatography (GC) In GC, the mobile phase which travels over the solid particles is a gas, usually something inert like helium or nitrogen. Instead of being dissolved in the gas, the mixture is usually just evaporated in the gas. The stationary phase consists of a solid powder which is stuck to the sides of the tube, very similar to TLC done in a tube. The difference here is that in GC, the column which is used to separate the compounds is extremely long (many meters, depending on how well you want it to separate the mixture). The principle is that if you have polar particles on the side of the tube and polar compounds evaporated in the gas, the polar compounds will be stuck to the particles more and evaporate less. In practice, this works extremely well for mixtures with VERY similar polarities. Often, GC machines have a mass spectrometer stuck to the other end of the nozzle. This gives you the ability to greater guess what your unknown is. We'll talk more about MS later.

Other exotic chromatographies Other chromatographic methods are also used, but usually not by high school students doing a lab. Here's a few:

High performance liquid chromatography (HPLC): HPLC is a method similar in principle to column chromatography, in which a liquid phase containing dissolved mixtures travels through a column containing a solid stationary phase. However, the difference here is that the columns are longer, and the pressures the liquid travels through the column are higher. For more information about HPLC, here's the website of a company that sells HPLC equipment: SciQuest. Reverse phase chromatography: Reverse phase chromatography is a form of chromatography where the solvent is polar and the stationary phase is nonpolar. I'll be honest with you: I don't know much about it, so I'm not going to spout off about it. Gel electrophoresis: Although it's not exactly a kind of chromatography, gel electrophoresis works under roughly the same principles. Basically, if you take a bunch of different nucleic acids, you can change the pH so they have

different charges. If you then apply an electric field across a gel in which you have placed the nucleic acids, the ones that have more charge will travel to one end of the plate faster than those that don't. This is pretty handy in biochemistry, where you need to be able to separate different strands of DNA and stuff like that. Size exclusion chromatography: Size exclusion chromatography is a form of chromatography where large particles in a mixture travel more slowly through a column because they can't get through as well as small ones. It's the same principle as in a traffic jam: Motorcycles can go as fast as they want in the worst traffic, because they can just go around other cars, since they're small. Trucks on the other hand, are stuck where they are because they're so big. The component in a mixture which travels fastest is the smallest one in SEC. 4. Spectroscopic methods What is spectroscopy? Spectroscopy is when you measure the strength of light in order to figure out what an unknown chemical is. There are two kinds of spectroscopy, absorption spectroscopy and emission spectroscopy. In absorption spectroscopy, you shine a light source at a sample containing your unknown and measure the light which passes through the sample. When you have these two values, you'll find that some of the light was absorbed by the sample - this energy corresponds to some energy level in the unknown. By knowing what energy level absorbed light, you can get clues about what the sample is. In emission spectroscopy, you add lots of energy to a sample and then measure the light that's given off. In one fairly common kind of emission spectroscopy, you burn up a sample in a very hot flame, and then measure the colors of light that are given off. These energies correspond to the differences in orbital energy levels in the element being studied. Because every element has different energy levels, you can identify an element by this manner. Let's look at some examples:

Infrared spectroscopy: Infrared spectroscopy (IR) is when infrared light is shone at a sample, and you measure how much of it is absorbed at each wavelength. As you might have guessed, this is a form of absorption spectroscopy. An aside: Infrared light is light that has a wavelength of between roughly 800 nm and ~20 microns, depending on who you talk to. This light is not visible to the naked eye... IR light is actually what we would refer to as being "heat". (If something is hot, it gives off lots of IR). In IR spectroscopy, the wavelength of light is transferred into units called "wavenumbers", where 1 wavenumber is equal to 1/cm (or put in other

ways, wavenumber is equal to the number of electromagnetic waves that can fit into one centimeter). It takes some getting used to. When you see an IR spectrum, it will have a series of broad and sharp peaks. Each of these peaks corresponds to some kind of functional group in the molecule. I'll let you look those up in an organic chemistry book somewhere rather than write them all out. So, what happens when this light is absorbed? Basically, the bonds in each molecule starts to vibrate, and the energy it takes to make the vibration occur accounts for the absorbed energy. IR spectroscopy is generally used for organic compounds, because it's good at identifying functionality. Because it doesn't give you an exact structure of your compound, you usually need some other information other than an IR spectrum to conclusively identify a compound. A cousin to IR spectroscopy is Raman spectroscopy. It works by roughly the same principles, although it's more involved. Suffice to say that if you're reading this, you're probably not about to jump right on a machine and start doing Raman spectroscopy.

UV-Vis (ultraviolet-visible) spectroscopy: UV-vis spectroscopy is a form of spectroscopy similar in principle to infrared spectroscopy. When you shine ultraviolet or visible light (the range for a spectrometer will usually go between about 300-1200 nm) on a sample, it will absorb some energy. The absorbed energy corresponds to some electronic transition in the molecule, such as an electron jumping from one orbital to another. As far as I can tell, every UV-vis spectrum is different. I haven't really seen much in the way of tables of functional group absorptions, mainly because every molecule is different. This method of spectroscopy is mainly handy for conclusively identifying an unknown - for example, if you know the unknown is one of two possibilities, and you know the UV-vis spectra for both, you just take a spectrum of the unknown and match it to the one you do know.

Nuclear magnetic resonance (NMR) spectroscopy: In NMR spectroscopy, radio waves are used to excite transitions in the molecules. Basically, you take a sample and place a small magnetic field on it. When this is done, the magnetic dipoles in the molecule all align in one direction. After this is done, you "pulse" the sample with an extremely strong radio wave, causing all the dipoles to align in the opposite direction. Eventually, all the dipoles go back to the way they were, causing energy to be given off. This energy is measured, and translated into a spectrum. For the spins to flip, you need atoms that have a spin = 1/2. Common atoms that do this are hydrogen-1 atoms (which are used in proton NMR), and carbon-13 atoms

(used in 13-C NMR). There are others. Many others. To put it plainly, NMR spectrometers are really cool. They look like giant metal balls with huge computer consoles attached. When a sample is inserted or removed, white smoke is given off by the boiling of liquefied nitrogen in the center. Why are they so elaborate? NMR spectrometers are basically huge magnets that give off enormous magnetic fields. These fields are required to cause the dipoles in molecules to switch direction. The liquid nitrogen is used to keep the magnets cool - because they are superconducting magnets, they need to stay at very cold temperatures. Do you want to really make an NMR spectroscoper mad? Just throw a handful of staples at the magnet. Then, watch the fun as they have to pry them off, one by one. Warning: If you throw staples at the magnet, the operator will cut your heart out and eat it. Warning: If you have a pacemaker, stay away from the magnet! Keep credit cards and disks away, too. NMR spectrometers are unbelievably expensive. Please, for your own safety, don't screw with the magnets or anything else, unless you want to lose your allowance for the rest of your life. By the way, MRI (magnetic resonance imaging) machines that they have in hospitals are really just big NMR spectrometers. Why the different name? Who knows? 5. X-ray crystallography (a.k.a. X-ray diffraction, or XRD) XRD is when you take x-rays and shine them at a crystal of a chemical. The stream of x-rays is then refracted by the crystal into a really complicated pattern. Using powerful computers, the XRD can tell you what molecule you have. XRD is mainly used for identifying the properties and structures of chemicals that you have already identified, although I suppose there's no reason that you couldn't use it to identify unknowns. Except for the reason that they're not all that common. And they're expensive. And they require you to have a crystal of your sample. Aside from that, no problem. Biochemists use XRD to identify what structures large proteins have. Other people use it, too, although I have no idea what for. In my years as a chemist, it never once came in handy. The main problem with XRD is that you need a crystal of your compound, and some compounds just don't crystallize. If your compound doesn't crystallize, then tough luck, buster, you ain't getting a diffraction pattern. 6. Mass spectrometry (a.k.a. mass spec or MS)

Mass spectrometry is really very handy for finding out what you've got in a sample. Essentially, it works like this: 1. Using a flame or laser, rip the molecules in your sample to shreds. Generally, this produces ions with positive charge. 2. Accelerate the ions toward a plate with negative charge. There should be a hole in the plate that the ions can pass through. 3. Steer the ions toward another plate with even more negative charge. If the particles have lots of mass, they'll travel further over the plate than particles with low mass because they have more momentum. The data from a mass spec isn't that hard to read. Basically, you end up with a chart that gives a list of peaks and corresponding m/z ratios. "m" stands for mass, and "z" stands for charge. (An example, if you have a particle that weighs 50 amu and a charge of +2, the m/z ratio will be 25). By interpreting this chart (or better yet, having a computer interpret the chart), you can figure out what the molecule looks like, and what elements are in it. Mass spectrometers can be added to the end of a gas chromatograph so that you can both separate mixtures and identify what the are.

Back to Mr. Guch's helpdesk! Back to the Cavalcade of Chemistry main page! Comments on this enormous essay? Email me at misterguch@chemfiesta.com

Nuclear Decay
Head back to the helpdesk Find more tutorials Try some practice worksheets

In this section :

Alpha decay Beta decay Gamma decay Positron emission Electron capture

As youve probably heard from a teacher, nuclear reactions are really important. Theyre important for people who are interested in blowing up

large parts of the world, theyre important for people who are interested in making huge quantities of energy without creating huge quantities of pollution, and theyre important if youre a doctor who wants to use radiation to treat various diseases. See, it really is important. In this help desk section, well talk about the different types of radioactive decay!

Alpha decay

Alpha decay occurs when helium nuclei come flying off of the nucleus of a larger isotope, forming an isotope with a smaller mass. These helium nuclei are called alpha particles, and are the same things that Rutherford busily shot at a sheet of gold foil during his experiment where he discovered the nucleus. When an atom undergoes alpha decay, the atomic number of the atom decreases by two and the atomic mass decreases by four. An example of alpha decay is shown below:

Beta decay
Beta decay is when an electron (called in this context a beta particle) is emitted from the nucleus of an atom, essentially turning a neutron into a proton. As a result, the atomic number of the element increases by one, while the mass stays virtually unchanged. An example of a beta decay is shown below:

Gamma decay

Gamma decay is when very high energy light called a gamma ray is emitted from a nucleus to bring it to a lower energy state. Gamma decay generally takes place at the same time as other nuclear reactions:

Positron emission

Positron emission is when a positron is given off by a nucleus. Positrons are the antimatter equivalent to electrons, so they have basically no mass and a charge of +1. Positron emission causes the atomic number of the element to decrease and the atomic mass to stay unchanged:

Electron capture

Electron capture is when an electron is absorbed by the nucleus of an atom, causing the atomic number to decrease by one and the atomic mass to stay unchanged. An example of an electron capture is shown below:

Questions, comments, interesting discoveries? Share them with me via email at misterguch@chemfiesta.com

http://misterguch.brinkster.net/explains2.html

Anda mungkin juga menyukai