Transcript for:
AP Chemistry Speed Review

Has AP Chemistry got you down? Does it seem like  TOO MUCH material? Hi there! My name is Jeremy   Krug, and in this AP Chemistry speed review,  we’re going to hit ALL the major topics in the   AP Chemistry course in less than 20 minutes.  Now this video can’t replace a full AP course,   but if you want to review, this is a good  place to start. But there’s a good chance you   might want an even better, more complete, maybe  even the ULTIMATE review packet, for AP Chem. I’m proud to be joining the AP content area  experts at Ultimate Review Packet dot com to   provide the absolute best AP Chem review  out there. If you’ve seen my YouTube review   videos and my full 101-video AP Chem course, then  you’re going to LOVE the Ultimate Review Packet,   because I’m in the process of preparing study  guides for every unit (and I’ll even give you   the answers), longer review videos that  cover tips and tricks for every unit,   plus a full length exam with answers. It’s  just $24.99 (and if you can get your teacher   to pay for the whole class, they’ll even  get a FORTY PERCENT DISCOUNT). So head over   to Ultimate Review Packet dot com and check it  out! Now, buckle up, here goes the Speed Review! Unit 1 covers atoms. We use a unit called  the MOLE to count large numbers of atoms   and molecules. When you weigh out a mole of  an element, it’s equal to its atomic mass,   or for a compound – the sum of those atomic  masses, expressed in grams. So one mole of iron   is about 55.85 grams. One mole of water,  H2O, is about 18.02 grams. Both of these   quantities have the same amount of fundamental  particles, 6.022 times 10 to the 23rd particles. You have to know electron configurations. For  neon, that’s 1s2, 2s2, 2p6. Atoms are usually   most stable when they have 8 electrons (or an  octet) in their outermost, or valence shell. We all know that opposite charges attract.  Well, Coulomb’s Law says that the greater   the magnitude of charge, the stronger  that attractive force will be. And the   closer together those charged particles  are, the stronger the attractive force   will be. This is why valence electrons  are held less tightly to the nucleus;   they’re literally farther away. And this is  how photoelectron spectroscopy works. Each   peak in this diagram represents a sublevel, and  the taller the peak, the more electrons it has.   The sublevels on the far left have electrons  that require more energy to be stripped away,   while the sublevels on the far right need way  less energy. So this diagram represents calcium. The Periodic Table has several patterns.  Like atomic radius: the bigger atoms are   at the bottom of the table, and also toward  the left. Atoms at the top and right of the   table have the highest first ionization  energy. When an atom gains electrons,   it becomes a negatively-charged anion, and  it gets larger. When an atom loses electrons,   it becomes a positively-charged  cation, and it gets smaller. Unit 2 covers chemical compounds. Metals and  nonmetals are held together by ionic bonds,   which are electrostatic forces: positives and  negatives attracting. Nonmetals are held together   by covalent bonds, where atoms share electrons.  Covalent bonds can be polar, where they share   electrons unequally, or they can be nonpolar,  where they’re sharing electrons fairly equally. Covalent bonds form molecules, which can  usually be individual units. However,   ionic compounds exist in a three-dimensional  lattice, where cations alternate with anions.   Metals, and metal alloys, exhibit metallic  bonding, where electrons can move freely,   consisting of positive metal ions  surrounded by a sea of electrons. Lewis electron-dot diagrams help us visualize  the shapes of molecules. Each atom has a certain   number of valence electrons, and we have to  draw every one of those in the diagram. There   are exceptions, but we try to arrange these  so every atom has eight valence electrons,   creating double or even triple bonds if we have  to. These shapes have names. This molecule has a   tetrahedral geometry and a bond angle of 109.5  degrees. This geometry would be called linear,   with a bond angle of 180 degrees, and  this one is trigonal planar, 120 degrees. Unit 3 covers Intermolecular Forces. Dispersion  forces are usually weak interactions,   but they get stronger as molecules get  larger and have more electrons. The more   electrons a molecule has, the more  polarizable it is. Dispersion forces   are the principal intermolecular  force between nonpolar molecules. Polar molecules also have dipole-dipole  forces. This is where the positive pole   of one molecule attracts the negative pole of a  neighboring molecule. These are usually stronger   than dispersion forces. Some polar molecules,  like water, have an especially strong force,   called hydrogen bonding, found between  molecules with an oxygen-hydrogen bond,   a nitrogen-hydrogen bond,  or a fluorine-hydrogen bond. Solids are usually crystalline, with  molecules packed tightly together,   having a fixed shape and volume. Liquids usually  have a little more space between the molecules,   so they can slip and slide around each other,  which is why a liquid flows. Gases have molecules   that are basically independent of each other. So  gases can expand or be compressed more easily. The Ideal Gas Law, PV=nRT, shows relationships  among pressure, volume, number of moles of gas,   and the temperature. However, the Ideal Gas Law is  just that, IDEAL. Gases in the real world aren’t   ideal, but they sometimes approximate ideal  conditions when we’re working with especially   small molecules, very weak attractions, or  even at a high temperature or low pressure. At higher temperatures, molecules have  a higher average kinetic energy. This   Maxwell-Boltzmann distribution shows that  at higher temperatures, more molecules are   moving faster, and at lower temperatures,  most molecules are moving more slowly. We measure solution concentration with  molarity, which is equal to moles of solute,   divided by liters of solution. The rule  “like dissolves like” helps us decide if   a solute will dissolve in a particular  solvent. Polar solutes dissolve in   polar solvents (like water). Nonpolar  solutes dissolve in nonpolar solvents. Light interacts with matter. The wavelength  of light, lambda, multiplied by it frequency,   nu, equals the speed of light. And if you  multiply that frequency by Planck’s constant,   you find the energy of a single photon of that  light. We often use spectrophotometry to analyze   the concentration of a solution. The higher the  absorbance of the solution by the instrument,   the higher the solution’s  concentration. We use this   data to build a graph to determine  the concentration of an unknown. Unit 4 covers chemical reactions. When we  write equations for reactions in solution,   we usually omit so-called ‘spectator’ ions that  don’t actually participate in the reaction, like   sodium or potassium cations, or nitrate anions.  The result is a net ionic equation. When you   write an equation, check that the number of atoms  of an element is the same on the reactant side as   it is on the product side. We use coefficients to  do this, and it’s called balancing the equation. A balanced equation is basically a recipe  for how the reaction works. The coefficients   in an equation form a MOLE RATIO. When  you make a stoichiometric calculation,   your first step should be to convert the quantity  to moles, if it’s not already in moles. Then,   use the coefficients of the balanced equation  to form a mole ratio to determine the moles of   the substance you’re converting  to. Finally, if you’re asked to   find a quantity in some other unit, like  grams, you’ll convert to that final unit. AP Chem focuses on three types of reactions. In  precipitation reactions, two solutions are mixed,   and a solid precipitate is formed.  In oxidation-reduction reactions,   one element loses electrons, a process  we call oxidation. At the same time,   another element gains those electrons, which  is called reduction. In a third type, acid-base   reactions, an acid reacts with a base to form a  conjugate acid and a conjugate base. Remember,   an acid is a proton donor, while a base is the  proton acceptor. Since a proton is just an H+ ion,   an acid always has exactly one  more H+ than its conjugate base. Unit 5 covers Kinetics. Balanced equations  help us describe relative rates. For example,   in this reaction, since the coefficient  of NH3 is twice that of N2, the rate of   appearance of ammonia will be twice as fast  as the rate of disappearance of nitrogen. Each reaction has its own rate law, and these are  determined experimentally. The rate law is always   written as: Rate equals k, which is the rate  constant, times the concentration of the first   reactant, raised to the power of its order,  times the next reactant, raised to its order,   and so on. If we double the concentration,  and the rate quadruples, that’s a second-order   relationship. On the other hand, if we double a  reactant’s concentration, and the rate doubles,   that’s a first-order relationship. If we double  the concentration, and the rate doesn’t change at   all, that’s zero-order. Once we know the order  for a reactant, we can use an integrated rate   law equation to calculate the amount that will be  left over after a certain amount of time. For each   integrated rate law, we have the rate constant k,  time elapsed t, initial concentration A subzero,   and elapsed concentration A sub T. If we know  any three values, we can calculate the fourth. Most reactions take place in multiple steps,  forming a reaction mechanism. One step is   slower than the others, and that slow step  determines the rate of the whole reaction. If   we can determine the rate law of the slow step,  we’ll know the rate law for the whole reaction. For molecules to react, they have to collide  with enough energy and in the right orientation.   When they do, a high-energy transition state can  form, which is at the peak on this graph. Then,   the products can be formed. The energy  required to start the reaction is called   the activation energy. This reaction has  a net loss of heat to the surroundings,   which makes it an exothermic reaction. To speed  up a reaction, we can raise the temperature,   use a smaller particle size, or raise  the concentration of the reactants. We   can also add a catalyst, which actually  provides a completely different reaction   mechanism and lowers the activation energy  required for the reaction to take place. Unit 6 covers Thermodynamics. Endothermic  reactions absorb heat from the surroundings,   while exothermic reactions release heat  into the surroundings. We calculate heat   transfer with the equation Q equals M C delta T. Q  represents heat in Joules, M is the mass in grams,   C is the specific heat capacity of the material,  and delta T is the change in temperature. The heat change for a reaction is called change in  enthalpy, or delta H, measured in kilojoules per   mole. We can estimate it using bond enthalpies,  adding all the enthalpies for the bonds broken   in a reaction, minus the total enthalpies for the  bonds formed. Or we can use enthalpy of formation,   where the sum of the enthalpies of formation  of the products, minus the total of the   enthalpies of formation of the reactants,  equals delta H. Or we can use Hess’s Law,   which says that if several individual reactions  add up to give a new reaction, the delta H values   of those individual reactions can be added  to give us the delta H of the new reaction. Unit 7 discusses equilibrium. When  a reaction reaches equilibrium,   it doesn’t stop. Instead, the forward reaction  has the same rate as the reverse reaction,   so the overall concentrations  of the substances stop changing. The expression for the reaction quotient is  abbreviated Q: the concentrations of the products   over the concentrations of the reactants,  raised to the power of the coefficients,   omitting any liquids or solids. If the reaction  is at equilibrium, then the reaction quotient   Q is equal to K, the equilibrium constant.  If we calculate Q and it’s not equal to K,   the reaction will proceed until it attains  equilibrium. When the equilibrium constant   K is very large, much greater than one, we’ll have  lots of product and very little reactant. However,   when K is very small, we’ll have lots  of reactant, and very little product. We’re often given initial concentrations  or pressures for a reaction and asked to   calculate the final concentrations or  pressures. The best way to do this is   to organize the data into a chart; I  call it an ICE box – initial, change,   equilibrium. Plug in the numbers you  know and solve for the numbers you   don’t know. Using algebra, you can  solve these problems like a puzzle. We can apply equilibrium in many ways. Le  Chatelier’s Principle says that if a reaction   is at equilibrium, and we add a component, the  reaction will shift toward the other side of   the reaction. And if we take away a product, the  reaction shifts so that product is replenished,   at the expense of the reactants. We can shift the  direction of the reaction, but the ONLY WAY to   change the actual value of the equilibrium  constant K is to change the temperature. Unit 8 is Acids and Bases. There are a few  essential equations in acid-base chemistry.   pH equals negative log of the hydrogen  ion concentration. pOH equals negative   log of the hydroxide ion concentration.  And at 25 degrees Celsius, pH plus pOH   of any solution equals 14. And at 25 degrees  Celsius, the hydrogen ion concentration times   the hydroxide ion concentration equals  1 times ten to the negative 14th power,   a constant we call Kw. Strong acids  and strong bases ionize completely,   so for example, in 0.50 molar nitric acid, the  concentration of the hydrogen ions is 0.50 molar,   so to find the pH you just  take negative log of 0.50. For weak acids and bases, less than 100  percent of the molecules dissociate,   so these are actually equilibrium  problems. Set up an ICE box,   where the initial concentration of the acid  or base is written in the appropriate spot,   the initial concentrations of the products  are zero, and you can solve the ICE box by   plugging into the equilibrium constant expression,  using Ka for a weak acid or Kb for a weak base. In acid-base titrations, we’re trying to find  the concentration of an acid or a base. We add   base to an acid until we reach the endpoint –  the moment where an indicator changes color,   signaling the reaction is complete. When  you plot pH versus volume of base added,   you get a titration curve that looks like  this. The inflection point represents the   equivalence point. And at pH 8.7, this  means the titration was between a strong   base and a weak acid. Halfway between the  starting point and the equivalence point,   the pH is 4.8, which tells us the Ka of  the acid is 10 to the negative 4.8 power. Buffers are mixtures of a weak acid and its  conjugate base that resist changes in pH.   We can calculate a buffer’s pH with  the Henderson-Hasselbalch equation. Unit 9 covers applications of thermodynamics.  Entropy, abbreviated S, is the disorder present   in matter. Solids have the least entropy,  since they are highly ordered. Pure liquids   have more entropy, solutions have even more,  and gases have the most entropy. Systems at   higher temperatures have more entropy than those  at lower temperatures. We can usually predict if   entropy is increasing or decreasing by looking  at a reaction. For example, when water melts,   solid is transitioning to liquid; entropy  is increasing, so delta S is positive. Gibbs Free Energy, delta G, is a measure of  the thermodynamic favorability of a process.   Delta G equals delta H minus temperature  in Kelvins times delta S. If a reaction   is thermodynamically favored at a certain  temperature, delta G is negative. If it’s NOT   favored, delta G is positive. Thermodynamic  favorability is related to the equilibrium   constant by the equation Delta G equals  negative R times T times natural log of K. In electrochemistry, every galvanic cell has two  half-reactions, one is an oxidation and one is a   reduction. The side where reduction  takes place is called the CATHODE,   and the side where oxidation takes place is  called the ANODE. Electrons move through the   wire from the anode to the cathode. The salt  bridge allows ions to flow freely through the   cell. At the salt bridge, anions flow toward  the anode, and cations flow toward the cathode. Every galvanic cell has a voltage that we can  calculate, using a list of standard reduction   potentials. As a galvanic cell runs, the voltage  slowly drops until it reaches zero volts, when   we say the cell is at equilibrium. We assume  galvanic cells are at standard conditions, which   is 25 degrees Celsius, and a concentration of one  molar for solutions. For any other conditions,   we use the Nernst Equation to calculate the actual  voltage. Galvanic cells are thermodynamically   favored, and its Delta G is equal to  negative n times Faraday’s Constant times E,   where n is the number of electrons transferred  and E is the overall voltage of the cell. An external electricity source passed through a  solution will power an electrolysis process. To   determine the amount of an element plated out, we  use this equation: Electrical current in amps, I,   equals electrical charge in Coulombs,  Q, divided by time in seconds. Once   you know the number of coulombs, you can  calculate the amount of metal plated out. There it is: The major points of the entire  AP Chemistry course in about 19 minutes.   Don’t forget to Like, Subscribe,  and watch my other review videos   and entire AP Chemistry Course.  If you’re taking AP Exams in May,   check out ultimate review packet dot com.  See you next time, and thanks for watching!