You are on page 1of 55

Chapter 7.

Chemical Energetics: Enthalpy


Faculty Resource and Organizational Guide (FROG)
Table of Contents
Materials for Chapter 7 Activities................................................................................................4
Reagents for Chapter 7 Activities: ...............................................................................................5
Section 7.1. Energy and Change ..................................................................................................6
Learning Objectives for Section 7.1 ............................................................................................6
Investigate This 7.1. What happens when foods are put in a flame?..........................................6
Consider This 7.2. Is energy involved when foods are burned?.................................................7
Consider This 7.3. How is release of thermal energy related to chemical bond energy?...........8
Section 7.2. Thermal Energy (Heat) and Mechanical Energy (Work).....................................9
Learning Objectives for Section 7.2 ............................................................................................9
Investigate This 7.4. What happens when a pinwheel is held over a flame?..............................9
Consider This 7.5. What causes a pinwheel over a flame to turn? ...........................................11
Consider This 7.7. How do thermal and mechanical energies interact?...................................12
Section 7.3. Thermal Energy (Heat) Transfer..........................................................................13
Learning Objectives for Section 7.3 ..........................................................................................13
Investigate This 7.8. Can radiation change the temperature of water?.....................................13
Consider This 7.9. How does radiation change the thermal energy of water? .........................15
Consider This 7.10. How does radiation transfer energy to water?..........................................16
Consider This 7.11. How are conduction and convection the same? different?.......................16
Section 7.4. State Functions and Path Functions .....................................................................17
Learning Objectives for Section 7.4 ..........................................................................................17
Investigate This 7.12. What happens when you rub your hands together?...............................17
Consider This 7.13. What change occurs when you rub your hands together? ........................18
Consider This 7.14. How do different pathways for descent of Pike's peak compare?............19
Consider This 7.15. How does the energy of the skin molecules change in Figure 7.5? .........19
Section 7.5. System and Surroundings......................................................................................20
Learning Objectives for Section 7.5 ..........................................................................................20
Consider This 7.18. How do systems and surroundings interact? ............................................20
Section 7.6. Calorimetry and Introduction to Enthalpy..........................................................22
Learning Objectives for Section 7.6 ..........................................................................................22
Investigate This 7.19. What happens when an acid and base are mixed? ................................22
Consider This 7.20. What causes the changes when an acid and base are mixed? ..................23
Investigate This 7.22. What happens when urea dissolves is water?........................................24
Consider This 7.23. What is T when urea dissolves in water? ...............................................25
Consider This 7.24. What is observed for an endothermic reaction in a calorimeter? .............25
Consider This 7.30. How can you get more accurate values for Hreaction?..............................26
Section 7.7. Bond Enthalpies......................................................................................................27
Learning Objectives for Section 7.7 ..........................................................................................27
Investigate This 7.32. What happens when yeast is added to hydrogen peroxide? ..................28
ACS Chemistry FROG

Chemical Energetics: Enthalpy

Chapter 7

Consider This 7.33. What reaction occurs when yeast is added to hydrogen peroxide?..........29
Investigate This 7.35. What is the gas formed in the hydrogen peroxide reaction? .................30
Consider This 7.36. Which bonds break and form in the hydrogen peroxide reaction?...........31
Consider This 7.38. How do bond enthalpies for single and multiple bonds compare?...........32
Section 7.8. Standard Enthalpies of Formation .......................................................................33
Learning Objectives for Section 7.8 ..........................................................................................33
Section 7.9. Harnessing Energy in Living Systems ..................................................................33
Learning Objectives for Section 7.9 ..........................................................................................33
Section 7.10. Pressure-Volume Work, Internal Energy and Enthalpy..................................33
Learning Objectives for Section 7.10 ........................................................................................33
Consider This 7.55 How is the first law of thermodynamics used? .........................................33
Consider This 7.58. How can E (or H) be the same for a reaction at constant volume and
constant pressure? ......................................................................................................................34
Section 7.11. What Enthalpy Doesn't Tell Us...........................................................................35
Learning Objectives for Section 7.11 ........................................................................................35
Section 7.13. Extension: Ideal Gases and Thermodynamics...................................................35
Learning Objectives for Section 7.13 ........................................................................................35
Investigate This 7.63. What happens when a gas is compressed or heated? ............................35
Consider This 7.64. What causes changes when a gas is compressed or heated? ....................36
Consider This 7.65. How are manometer readings interpreted and quantified?.......................37
Consider This 7.66. How does kinetic-molecular theory apply to Investigate This 7.63? .......38
Investigate This 7.72. Do reactions in open containers and capped containers differ? ............39
Consider This 7.73. Why do reactions in open and capped containers give different results? 43
Consider This 7.74. What are qV and qP for the hydrogen carbonate-acid reaction? ...............44
Consider This 7.76. What is w for the hydrogen carbonate-acid reaction? ..............................44
Solutions for Chapter 7 Check This Activities...........................................................................46
Check This 7.6. The difference between warm and cool gases ................................................46
Check This 7.16. The system in Figure 7.5 ..............................................................................46
Check This 7.17. Closed and open systems..............................................................................46
Check This 7.21. Is the reaction of an acid with a base exothermic or endothermic?..............46
Check This 7.25. Energy transfers in a calorimeter..................................................................46
Check This 7.27. Determination of qP(reaction) for dissolution of urea in a calorimeter .............47
Check This 7.29. Determination of Hreaction for dissolution of urea in a calorimeter .............47
Check This 7.31. Effect of increasing the amount of another calorimetric reaction ................47
Check This 7.34 Properties of the possible H2O2 reaction products.........................................47
Check This 7.37. Overall homolytic bond cleavage reaction of methane ................................47
Check This 7.39. Formation of polyethylene............................................................................48
Check This 7.40. Sulfursulfur bonds ......................................................................................48
Check This 7.42. Reaction enthalpy from average bond enthalpies.........................................49
Check This 7.43. The hydrogen peroxide decomposition reaction ..........................................51
Check This 7.45. Formation of compounds from their standard-state elements ......................51
Check This 7.47. Standard enthalpy change for an isomerization reaction..............................51
Check This 7.48. Comparison of formation of 1-butene from its atoms and its elements .......52
Check This 7.50 .........................................................................................................................52
2

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

Check This 7.51 .........................................................................................................................53


Check This 7.52 .........................................................................................................................53
Check This 7.53 .........................................................................................................................54
Check This 7.54 .........................................................................................................................54
Check This 7.57 .........................................................................................................................54
Check This 7.60 .........................................................................................................................54
Check This 7.62 .........................................................................................................................54
Check This 7.67 .........................................................................................................................54
Check This 7.69 .........................................................................................................................55
Check This 7.71 .........................................................................................................................55
Check This 7.77 .........................................................................................................................55

ACS Chemistry FROG

Chemical Energetics: Enthalpy

Chapter 7

Materials for Chapter 7 Activities


Activity

Material

Total Quantity

7.1, 7.4

Bunsen burner or candle

7.1, 7.4

Matches or lighter

7.1

Tongs

7.1

Pan of water

7.4

Long metal cylinder

7.4

Plastic or paper pinwheel

7.8

Bright incandescent bulb

7.8

Lamp

7.8

Narrow flat-sided clear


colorless containers

7.8

Reflector

7.8

Thermometers or
temperature probes

7.19

Small test tubes

7.19

Plastic pipets

7.22

Calorimeter; Styrofoam
cup, cover, and temperature
probe or thermometer

7.32

100 mL graduated cylinder

7.35

Long fireplace match or


wooden splint

7.35

Stirring rod

7.36

Molecular model kits

1/student

7.63

50-mL plastic syringe

7.63

Plastic tubing

7.72

2-L soft drink bottle

7.72

Watch or timer

7.72

50-mL plastic syringe

7.72

New 24/40 ribbed rubber


septum

7.72

Powder funnel

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

Reagents for Chapter 7 Activities:


Activity
7.1

Reagents
Spaghetti, minimarshmallow, potato chip

Total Quantity
1-3 of each.

Optional: Nuts (such as


walnuts, peanuts, and/or
pecans), corn chip, piece of
chocolate
7.19

1M NaOH

10 mL

7.19

1M HCl

10 mL

7.22

Urea, H2NC(O)NH2(s)

6.0 g

7.32

3% hydrogen peroxide
solution, H2O2

10 mL

7.32

Liquid dishwashing
detergent

1-2 mL

7.32

Yeast

2 to 3 g

7.63

Food coloring

several drops

7.72

Sodium hydrogen carbonate


(baking soda, NaHCO3(s))

2 8.4 g

7.72

6 M hydrochloric acid
(HCl) solution.

2 20.0 mL

ACS Chemistry FROG

Chemical Energetics: Enthalpy

Chapter 7

Section 7.1. Energy and Change


Learning Objectives for Section 7.1
Identify some forms of energy observed in chemical changes and learn that energy is
conserved in the changes.
Investigate This 7.1. What happens when foods are put in a flame?
Goal:
Observe that energy is released (heat and light) when foods are burned and get a qualitative idea
of the energy content of various foods.
Set-up time:
Approximately 15 minutes (assuming food products are already available).
Time for activity:
Approximately 5-10 minutes (depending on how many food products you test).
SAFETY NOTE
Wear your safety goggles
If someone in the class has a food allergy to one of the
food items you choose, do not burn that item.
Equipment:
Bunsen burner or candle.
Matches or lighter.
Tongs (optional).
Pan of water.
Reagents (foods):
Spaghetti.
Mini-marshmallow.
Potato chip.
Optional: Nuts (such as walnuts, peanuts, and/or pecans), corn chip, piece of chocolate.
Procedure:
This activity is conducted as a class activity.
Ask for a student volunteer and provide safety goggles.
Have the student take a full-length spaghetti piece and hold it by one end over the pan of
water.
Bring the flame under the free end of the spaghetti and hold it there until its appearance begins
to change. Then remove it from the flame. If necessary, have the student extinguish the
spaghetti by immersing it in water.
The rest of the class should record their observations.
Repeat with a marshmallow (or other food product) held on the end of another full-length
piece of spaghetti.
Instead of using a full-length piece of spaghetti as a holder, tongs may be used.
Clean-up:
All fully extinguished food products should be disposed in a trash container.
Anticipated results:
Here are photographs of a few foods burning:

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

spaghetti

mini-marshmallow

pecan

potato chip
corn chip
chocolate
Follow-up discussion:
Use Consider This 7.2 and 7.3 to initiate and facilitate classroom discussion of this activity.
Discuss this activity qualitatively food as a source of energy, conversion of one form of
energy into another, etc.
The discussion should introduce students to the first law of thermodynamics as well as the
vocabulary used in measuring energy.
Follow-up activities:
End of chapter problems 7.1 through 7.3.
Consider This 7.2. Is energy involved when foods are burned?
Goal:
Students should conclude that energy is released differently when different foods are burned and
relate this observation qualitatively to the composition of various foods.
Classroom options:
Allow 3-5 minutes for students, working in small groups, to answer these questions. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
This activity can be conducted as an open class discussion.
Time for activity:
Approximately 10 minutes.
Instructor's notes:
This activity is a follow-up discussion to Investigate This 7.1 and serves to focus students on
the idea that energy is released when food is burned and that this is the energy that is provided
by our metabolic processes.

ACS Chemistry FROG

Chemical Energetics: Enthalpy

Chapter 7

When the foods burn in this activity, they combine with oxygen to produce mostly carbon
dioxide and water.
Stress that the form of energy changes, but energy is always conserved as stated in the first
law of thermodynamics.
Students should reason and conclude:
(a) When the foods were placed in the flame some of them, such as marshmallows, chips,
and nuts, ignited and burned, even when removed from the flame. Others, such as spaghetti,
were less able to sustain a flame, but continued to glow and char when out of the flame. A
flame was required to ignite the foods and light and heat were released when the foods
burned on their own; these observations are evidence that energy was involved in the
changes. All of the products shown in the chapter opener can be ignited and will burn or
char, as we have observed for a few of them. The flames of each of the burning foods was
different. Yes, once the energy in these foods is metabolized by our bodies, it provides us
with energy.
(b) The foods did not all burn the same way, as noted in part (a). Chips and marshmallows
burn quickly. Nuts burn steadily. Spaghetti burns slowly and not as vigorously. Chips contain
a good deal of fat as well as starch and the fat burns quickly. (Do baked chips burn the same
as fried chips?) Marshmallows are mainly sugar and gelatin (protein) and are puffed up,
giving them lots of surface area for reaction with oxygen from the air. Nuts are full of oil,
which burns well but has to ooze to the surface of the nut, in order to vaporize and burn, so
nuts dont usually burn rapidly. Spaghetti is mostly starch (a polymer of glucose), very
similar to the cellulose in wood, and it burns much like wood. Perhaps these observations on
the foods are somewhat indicative of their energy release over time, but metabolic processes
are different than combustion, so we should be careful not to draw too many conclusions
about metabolism from these observations.
Follow-up activities:
Consider This 7.3. How is release of thermal energy related to chemical bond energy?
End of chapter problems 7.1 through 7.3.
Consider This 7.3. How is release of thermal energy related to chemical bond energy?
Goal:
Students should link the energy release observed in Investigate This 7.1 to the relative strengths
of the chemical bonds in the products and reactants.
Classroom options:
Allow 3-5 minutes for students, working in small groups, to answer these questions. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
This activity can be conducted as an open class discussion.
Time for activity:
Approximately 10 minutes.
Instructor notes:
Review the results from Investigate This 7.1 to be sure the class agrees on the observations.
Discuss the fact that different compounds are being burned in each food, but that the resulting
products are pretty much the same in each sample (carbon dioxide + water). [Of course there
are other products produced by the atoms other than carbon and hydrogen in the foods, but
they do not affect the argument about relative bond strengths.]

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

Students should reason and conclude:


All the foods burned and released energy in the form of light and heat. Thus, the change from
the chemical bonds in the reactants, oxygen and the compounds in the food, to the chemical
bonds in the products, mostly carbon dioxide and water, releases energy. The bonds in the
products must be stronger, that is, lower in energy than those in the reactants, in order that
the change should release energy. Because we have no way to know how much fuel was
burned in each case, we cannot compare the relative energy output of the foods from our
qualitative observations in Investigate This 7.1.
Follow-up activities:
End of chapter problems 7.1 through 7.3.

Section 7.2. Thermal Energy (Heat) and Mechanical Energy (Work)


Learning Objectives for Section 7.2
Identify the forms of energy transferred in physical and chemical changes and show how we
can interpret energy conservation in the changes.
Draw molecular level representations of thermal energy (undirected kinetic energy) and
mechanical energy (directed kinetic energy) transfers.
Investigate This 7.4. What happens when a pinwheel is held over a flame?
Goal:
Observe the motion of a pinwheel placed over a burner with and without a chimney.
Set-up time:
At least 20 minutes initially. This activity is dependent on type of pinwheel, type of chimney,
and room ventilation. Practice this activity with the pinwheel, heat source, and chimney you
will use before doing it in your class.
Time for activity:
Less than 10 minutes.
Equipment:
Bunsen burner or microburner. A candle or an incandescent electric light bulb will work, if
the pinwheel is very lightweight and has little resistance to turning.
Long metal cylinder. A 3-foot long aluminum conduit tube with a 3.5 cm inner diameter
seems to work with a microburner (see photograph below). A metal tube vacuum cleaner
attachment can be used. A glass cylinder could also be used. Soup cans (or similar shaped
cans) with both ends cut out can be taped together (duct tape) to make a cylinder, but not
everyone has success with this chimney.
Metal, plastic or paper pinwheel. Commercial toy pinwheels generally are too heavy and exert
too much friction on their bearing surface to turn well when held "face down" over a flame.
Paper pinwheels work well.
NOTE: Here are two recipes for constructing paper pinwheels.
Construct a simple pinwheel by cutting out a 3-inch circle from a 3 by 5 card, cutting slits
about an inch long along the radii at 60 intervals, bending one edge of each slit into a fin, and
suspending the pinwheel on a length of thread passing through a tiny hole in the center of the
pinwheel and knotted at one end to prevent the thread going through the hole. This is the set
up shown in the photograph accompanying Investigate This 7.4 in the textbook. When doing
the investigation, allow the pinwheel to hang from its support, a ring on a tall ring stand, for
example, until it stops rotating (due to twisting of the thread). When the source of heat is
ACS Chemistry FROG

Chemical Energetics: Enthalpy

Chapter 7

moved below it, the pinwheel should turn. Without a chimney to direct the flow of hot gases,
the motion will probably be slow, but should speed up when the chimney is in place. The
thread will tend to wind up, so that the motion slows over time and, when the source of heat is
removed, the pinwheel reverses its rotation as the thread unwinds. The candle shown in the
photo can be replaced with a 75- or 100-watt incandescent light bulb as the source of heat for
gases in the air and students might be asked to draw a figure analogous to Figure 7.2 to
explain what is going on in this case.
Construct a paper pinwheel using these materials: strong construction paper, long dressmaker
pin, penny, and a pencil with an eraser or a thin wood dowel. Cut the paper into a 6" by 4"
rectangle. Draw two diagonal lines, corner to corner. Trace a circle using the penny in the
center where the diagonal lines cross. Remove the penny and then cut along the diagonals
from each corner of the paper to the edge of the circle made with the penny. Fold (without
creasing) alternate corners into the center and fasten together with the pin. Stick the pin firmly
into the top of the dowel or the eraser of the pencil. Bend the pin so the pinwheel is parallel to
the pencil or dowel (acting as a handle). These directions are adapted from Gray, Judith, and
Ellison, Sheila, 365 Days of Creative Play: For Children 2 years and Up, Sourcebooks, Inc.
Procedure:
Due to the experimental variability of this activity, you should conduct it yourself in class,
after practicing with your apparatus. The procedure described below uses a microburner, a
paper pinwheel, and an aluminum metal conduit tube.
SAFETY NOTE
Wear your safety goggles.
Be careful not to burn yourself or the pinwheel with the
burner flames or other heat sources.
Set the burner to a low, flickering yellow flame.
Hold your hand about 25 cm above the flame and note how warm the combustion gases
coming from the flame fee. Keep raising your hand until you cannot feel the heat from the
flame, noting the distance between your hand and the flame. Note: We could feel the heat up
to 2 ft above the microburner.
Hold the pinwheel about 25 cm above the flame and observe its motion, if any. Try different
orientations and angles to find the one that causes the pinwheel to spin the fastest. Note: Our
paper pinwheel spun slowly at a slight angle about 3 ft above the bench top slightly off center
of the burner.
Place the metal cylinder over the burner so that it serves as a chimney. Wait a few moments
for the metal to become warm.
Hold your hand just above the top of the chimney and note how warm the combustion gases
from the flame feel. Move your hand upward until you do not feel any more heat. Compare
your results to holding your hand over the burner at the same distance without the chimney.
Note: We could feel heat 5 feet above the burner.
Hold the pinwheel just above the top of the chimney and observe its motion, if any. Try
different orientations to find the one that causes the pinwheel to spin the fastest. Note: We
observed fast pinwheel spinning even at 2 feet above this type of chimney. Better results are
obtained with a longer, narrower tube.
If your pinwheel does not spin, check to make sure there is no friction between the pinwheel
and its axle.

10

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

Anticipated results:
This is a photograph of the set up described above.

Follow-up discussion:
Use Consider This 7.5 to initiate discussion of this activity.
Follow-up activities:
Check This 7.6. The difference between warm and cool gases.
Consider This 7.7. How do thermal and mechanical energies interact?
End of Chapter problems 7.4 through 7.8.
Consider This 7.5. What causes a pinwheel over a flame to turn?
Goal:
Discussion of the observations from Investigate This 7.4 should lead students to conclude that
the movement of hot gas can be directed (by the chimney) to perform work (cause a pinwheel to
spin).
Classroom options:
Allow 3-5 minutes for students, working in small groups, to answer these questions. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
This activity can be conducted as an open class discussion.
Time for activity:
Approximately 10-15 minutes.
Instructor notes:
Use this discussion to introduce students to thermal and mechanical energy as two different
forms of kinetic energy.
The discussion should also elucidate the function of the chimney in Investigate This 7.4.
Without a chimney, the gases leaving the flame dissipate. Not enough of the hot gas goes
directly upward to spin the pinwheel. With the chimney, a greater upward-directed flow of hot
gas occurs and turns the pinwheel. The temperature difference noted with and without the
chimney is convincing evidence that more hot gas is directed upward with the chimney.
Discuss the difference between directed and undirected kinetic energy and that both thermal
energy and mechanical energy are forms of kinetic energy.
Figure 7.2 can be used to help explain the microscopic movement of gases in this activity.
Students should reason and conclude:
ACS Chemistry FROG

11

Chemical Energetics: Enthalpy

Chapter 7

(a) A difference in temperature of the combustion gases without and with a chimney is
observed. The combustion gases with the chimney in place felt warmer than at the same
distance above the burner without the chimney. More of the hot gas from the combustion is
directed upward by the chimney. The result is that the gas feels warmer at a given distance
above the flame, because there is more hot gas to transfer energy to your hand.
(b) The pinwheel turned notably faster with the chimney placed over the burner. This must
mean that a larger amount of gas (more molecules) was hitting the pinwheel when the
chimney was in place.
(c) The gas at the outlet of the chimney felt warmer and the pinwheel in this position spun
faster. In part (a) we concluded that more hot gas was reaching our hand when the chimney
was in place to direct the flow. In part (b) we concluded that the pinwheel spun faster above
the chimney because more gas molecules were hitting it. Our correlation between warmer gas
and faster pinwheel spin is consistent with our conclusion (explanation) that more hot gas is
directed upward by the chimney.
Follow-up activities:
Check This 7.6. The difference between warm and cool gases.
Consider This 7.7. How do thermal and mechanical energies interact?
End of chapter problems 7.4 through 7.8.
Consider This 7.7. How do thermal and mechanical energies interact?
Goal:
Conclude that thermal and mechanical energies can be interchanged.
Classroom options:
Allow 3-4 minutes for students, working in small groups, to answer these questions. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
Time for activity:
Approximately 10-15 minutes.
Instructor notes:
Use the discussion to reinforce the idea that energy is conserved but the form of energy may
change.
Try to be sure that students link the macroscopic level of a pinwheel spinning with the
microscopic motions of the hot gases that strike it to make it turn.
Students should reason and conclude:
To make the pinwheel spin, the rising gases have to collide with it and give up some of their
energy of motion (kinetic energy) to the pinwheel vanes. Thus, the gas molecules that have
collided with the pinwheel will, on the average, be moving more slowly after the collision.
Slower moving gas has less kinetic energy and, hence, a lower temperature. Collisions with
the pinwheel cool the rising gas more than if the pinwheel was not present.
Follow-up activities:
End of chapter problems 7.4 through 7.8.

12

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

Section 7.3. Thermal Energy (Heat) Transfer


Learning Objectives for Section 7.3
Identify the forms of energy transferred in physical and chemical changes and show how we
can interpret energy conservation in the changes.
Draw molecular level representations of thermal energy (undirected kinetic energy) and
mechanical energy (directed kinetic energy) transfers.
Identify whether a thermal energy transfer occurs by radiation, conduction, and/or convection.
Investigate This 7.8. Can radiation change the temperature of water?
Goal:
Observe the temperature of water in containers with a bright light shining on them.
Set-up time:
Approximately 10 minutes, if the light bulb and reflector set up are available.
Time for activity:
From about 10 minutes to 20-30 minutes. (Some instructors have the class monitor the
temperatures for the entire class session.)
Equipment:
Bright incandescent bulb (either 100, 150 or 200 watt). The light beam has to be unfiltered, so
that plenty of infrared radiation is included. Light from a slide projector is usually filtered to
prevent overheating the slides, so this is usually not a good source.
Lamp holder (a lamp kit can be used as shown in the set-up below).
2 narrow flat-sided clear, colorless containers filled with room temperature water. Plastic, 75cm2 cell-culture flasks used in microbiology work very well.
Reflector (one can be made by covering poster board with aluminum foil as shown in the setup below).
Two standard laboratory thermometers. Digital thermometers or temperature probes
interfaced to a computer with the results projected for the whole class to observe are excellent.
Procedure:
Conduct this investigation as a class activity, using students to read the temperatures (unless
computer interfaced probes with projected output are used).
SAFETY NOTE
Wear your safety goggles.
The set-up is shown here:

ACS Chemistry FROG

13

Chemical Energetics: Enthalpy

Chapter 7

In the set-up shown, the thermometer in the left bottle was placed 7 cm from the surface of the
light bulb and the thermometer in the right bottle 12 cm from the surface of the light bulb.
Record the temperature of the water in the containers.
Direct a bright, incandescent light beam through the containers.
After 10 minutes in the light beam, read and record the temperature in the two containers.
Remove the container nearer the light source and read and record the temperature in the
remaining container after another 10 minutes.
NOTE: The outcome of this simple activity is not obvious to students who have never
experienced the effect of infrared absorption by water (or never thought about the origin of the
heat absorption by water). The setup is designed to minimize conduction and convection effects
and students are asked to think about this as well.
Anticipated results:
The results for three different conditions are given here.
100-watt bulb:

Initial reading
10 minutes

Temperature in Nearer Bottle

Temperature in Further Bottle

21.5C

22C

25C

22.5C

20 minutes

24C

100-watt bulb;
Temperature in Nearer Bottle

Temperature in Further Bottle

Initial reading

22C

21.7C

15 minutes

27C

23C

30 minutes

25.5C

150-watt bulb:
The temperature of the nearer container after 10 minutes is about 10 C warmer than at
the start, but the further container only 2-3 degrees warmer. When the nearer container is
removed, the water in the remaining container will be almost as warm (after 10 more
minutes) as the nearer container before it was removed.
Instructor notes:
Use the waiting time in this activity to discuss temperature, thermal energy, and heat.
Temperature is a measure of the average kinetic energy of the particles in a substance.
Thermal energy is the sum of all the kinetic energy of the particles. For a given object or
amount of material, an increase in thermal energy will increase the temperature (because the
average particle is moving faster) and vice versa.
Heat is a measure of the thermal energy transferred between two objects at different
temperatures (transfer from hot to cold).
Follow-up discussion:
Use Consider This 7.9 to initiate discussion of the observations in this activity and try to get to
Consider This 7.10 and 7.11 while the experimental set-up is still available for testing.

14

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

Follow-up activities:
Consider This 7.10. How does radiation transfer energy to water?
Consider This 7.11. How are conduction and convection the same? Different?
End of chapter problems 7.9 through 7.11.
Consider This 7.9. How does radiation change the thermal energy of water?
Goal:
Students should conclude that radiation can be absorbed by and change the thermal energy of
water.
Classroom options:
Allow 3-4 minutes for students, working in small groups, to answer these questions. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
This activity can be conducted as an open class discussion.
Time for activity:
Approximately 10 minutes.
Instructor notes:
Try to be sure that the class understands the relationship of thermal energy and temperature of
a sample of water as this discussion is carried on.
Try to get to Consider This 7.10 and 7.11 while the experimental set-up from Investigate This
7.8 is still available for testing.
Students should reason and conclude:
(a) The temperatures in both containers increased. The temperature of the water in the
container nearer the light bulb increased more than the temperature of the water in the one
behind it. Since temperature is a measure of thermal energy and the temperature of the water
in both containers increased, the thermal energy of the water increased in both containers.
The thermal energy in the container nearer the light source increased more, as evidenced by
its greater increase in temperature.
(b) After 10 minutes in the absence of the nearer container, the temperature of the water in
the remaining container increased to almost the same temperature as in the nearer container
when it was removed. Since temperature further increased, we conclude that the thermal
energy of the water has increased (to almost the same amount as in the container that was
removed). If radiation (light energy) is absorbed by water to increase its thermal energy and
hence its temperature, this would explain the increase in temperature of the water in the
containers. Since some of the radiant energy is absorbed by the sample of water in the nearer
container (in order to increase its thermal energy), this energy will be missing from the
radiation that goes into the second container. If there is not as much energy to be absorbed,
the water in the second container will not gain as much thermal energy and its temperature
will not increase as much as that in the first, as we observed. When the nearer container is
removed, all the radiant energy can reach the second container, the water will gain more
energy, and the temperature will increase, as observed. [Since the light beam is not really
focused through the samples, the container farther from the source will not receive quite as
much radiant energy and will not warm quite as much in the same time as the nearer one.]
Follow-up Activities:
Consider This 7.10. How does radiation transfer energy to water?
Consider This 7.11. How are conduction and convection the same? different?

ACS Chemistry FROG

15

Chemical Energetics: Enthalpy

Chapter 7

End of chapter problems 7.9 through 7.11.


Consider This 7.10. How does radiation transfer energy to water?
Goal:
Conclude, based on their observations and data from Investigate This 7.8 and the discussion that
precedes this activity, that infrared radiation is absorbed by water and increases its thermal
energy.
Classroom options:
This activity can be an extension of the discussion begun with Consider This 7.9.
Time for activity:
Approximately 10 minutes.
Instructor notes:
Remind students of the terms used to explain the concepts behind these activities. Chapter 4
may need to be reviewed, especially electromagnetic radiation wavelengths and energy
emission by glowing bodies.
If possible, introduce conduction and convection as a follow-up to this activity while the setup for Investigate This 7.8 is still available to be analyzed. By all means have some students
try the experiment suggested at the end of part (b) of this activity. This is quite a good
demonstration of the infrared absorbing property of water.
Students should reason and conclude:
(a) Water molecules absorb infrared radiation from the light beam, which causes them to
move faster, that is, have greater thermal energy. We detect this increase as an increase in the
temperature of the water sample. We cannot see any difference in the light beam before and
after it enters the water, because infrared radiation is invisible to our eyes (see Figure 4.13,
page 227).
(b) You would feel a difference in your hand placed in the light beam before and after it
passes through a container of water. When you place your hand in the light beam before it
passed through the container of water, your hand will absorb the infrared radiation and you
will feel your hand getting warmer. When you place your hand in the beam after it passes
through the water, you will notice that your hand gets less warm because the water molecules
in the container will have absorbed much of the infrared radiation from the light beam.
Follow-up discussion:
Use the discussion of radiation energy transfer to introduce conduction and convection as
modes of energy transfer that require the presence of matter.
Follow-up activities:
Consider This 7.11. How are conduction and convection the same? Different?
End of chapter problems 7.9 through 7.11.
Consider This 7.11. How are conduction and convection the same? different?
Goal:
Students should conclude that conduction requires particles/bodies in contact and convection
involves the movement of particles from one place to another, and should apply these ideas to
the Investigate This 7.8 set-up.
Classroom options:
This activity can be an extension of the discussion begun with Consider This 7.9 and
continued with Consider This 7.10.

16

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

Time for activity:


Approximately 10-15 minutes.
Instructor notes:
Show or refer students to Figure 7.3 as this activity is carried out.
Students should reason and conclude:
(a) The fundamental difference between conduction and convection is that the particles
originally energized by the thermal energy source are often the ones we sense in convection.
That is, the particles move from the source to the sensor. But, in conduction, there is a relay
or domino effect from the originally-energized particles through many intermediate particles
to the sensor. The particles do not move from the source to the sensor. The mass of particle
moving in a convection current can, of course, pass on their energy to other particles when
they collide with them, so some of particles that reach the sensor have been energized by
collisions along the way, not directly by the energy source.
(b) There could be some minor conduction and/or convection effects in Investigate This 7.8.
Convection would require motion of the air from the source to the containers of water. Since
the containers are not above the energy source, there is probably not very much convective
heating, as long as the light bulb apparatus is pretty well open so that heated air does not
contact the containers. Air (and other gases) is not a very good conductor of thermal energy,
because the molecules are far apart and transfer of energy between them is not nearly as rapid
as in a liquid or solid. A gentle flow of air between the light and the containers would make
sure that warmed air molecules were swept away before they could reach the containers of
water. If the results are still the same, with this intervention, then we can be more certain that
convection and conduction are not responsible for the observed temperature changes in the
investigation.
Follow-up activities:
End of chapter problems 7.9 through 7.11.

Section 7.4. State Functions and Path Functions


Learning Objectives for Section 7.4
Identify the forms of energy transferred in physical and chemical changes and show how
energy is conserved in the changes.
Define and identify the variables that are functions of state and those that are functions of the
path for a given change.
Investigate This 7.12. What happens when you rub your hands together?
Goal:
Conclude that rubbing your hands together makes the palms feel warm.
Set-up time:
None.
Time for activity:
1-2 minutes.
Procedure:
Students place their palms of their hands together and rub them briskly for 3-4 seconds.
Follow-up discussion:
Use Consider This 7.13 to initiate discussion of the results of this activity.

ACS Chemistry FROG

17

Chemical Energetics: Enthalpy

Chapter 7

Follow-up activities:
Consider This 7.14--How do different pathways for descent of Pike's Peak compare?
Consider This 7.15--How does the energy of the skin molecules change in Figure 7.5?
End of Chapter problems -- 7.12-7.16.
Consider This 7.13. What change occurs when you rub your hands together?
Goal:
Students should conclude that mechanical motion can be converted to thermal energy and the
same change in thermal energy can be brought about in another way, thus introducing the
concepts of path function and state functions.
Classroom options:
Allow 3-4 minutes for students, working in small groups, to answer these questions. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
This activity could be conducted as an open class discussion.
Time for activity:
About 5-10 minutes.
Instructor notes:
Use this discussion as an introduction to the importance of defining thermodynamic variables
such as temperature, thermal energy, and mechanical energy.
Students should begin to understand the variables involved with thermodynamics as well as
the difference between state functions and path functions.
Students should reason and conclude:
(a) Because temperature is a measure of the average motion of molecules, the molecules in
your skin must have been made to move faster by rubbing your hands together, since your
hands felt warmer after the rubbing.
(b) Mechanical energy supplied by your muscles to rub your hands together is transferred to
molecules at the surface of the skin (simply by pushing on them), thus increasing the kinetic
energy of the skin molecules, which we sense as an increase in temperature (larger amount of
thermal energy in the skin due to the rubbing).
(c) Yes, you could use a heater, a candle, or a campfire to warm your hands. Also, a pair of
mittens can be worn. Thermal energy that is continuously released by metabolic processes in
our bodies is trapped in the fibers and air pockets of the mittens. Since the mittens are then at
a higher temperature than the surrounding air, less thermal energy is transferred from the skin
(because the transfer depends on the difference in temperature between the object that is
releasing the energy and the body that is receiving it) and the thermal energy builds up in the
skin, making your hands inside the mittens feel warmer.
Follow-up discussion:
Show or refer students to Figures 7.4 and 7.5 during this discussion, in order to reinforce that
thermal and mechanical energy transfers are path functions.
Follow-up activities:
Consider This 7.14. How do different pathways for descent of Pike's Peak compare?
Consider This 7.15. How does the energy of the skin molecules change in Figure 7.5?
End of chapter problems 7.12 through 7.16.

18

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

Consider This 7.14. How do different pathways for descent of Pike's peak compare?
Goal:
Conclude that the distance traveled on the ground for the same change in elevation is different
for different pathways.
Classroom options:
Allow 2-3 minutes for students, working in small groups, to answer these questions. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
This activity could be conducted as an open class discussion.
This activity could also be assigned as a homework problem and then discussed at the next
class session.
Time for activity:
About 5-10 minutes.
Instructor notes:
Display Figure 7.4 (Two ways to get to the top of Pike's Peak) before conducting this activity.
Students should reason and conclude:
Even though your change in elevation (a state function) is the same by both pathways, your
car's odometer will indicate that you traveled more linear ground distance than your friend,
because you had to use a winding mountain road as compared to the more direct descent via
the cog railway.
Follow-up discussion:
Use this discussion to help define and distinguish path functions from state functions, linking
the discussion to Investigate This 7.12, Consider This 7.13 and Figure 7.5.
Follow-up activities:
Consider This 7.15. How does the energy of the skin molecules change in Figure 7.5?
End of chapter problems 7.12 through 7.16.
Consider This 7.15. How does the energy of the skin molecules change in Figure 7.5?
Goal:
Conclude that different paths for warming hands produce the same energy change in the hands.
Classroom options:
This activity can be conducted as an open class discussion.
Allow 2-3 minutes for students, working in small groups, to answer these questions. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
This activity could also be assigned as a homework problem and then discussed at the next
class session.
Time for activity:
Approximately 5-10 minutes.
Instructor notes:
Show or refer students to Figure 7.5 during the discussion in this activity.
Students should reason and conclude:
Although the paths are different, the energy change for the skin molecules is the same
because the temperature change, a measure of the thermal energy change, is the same.
Follow-up activities:
End of chapter problems 7.12 through 7.16.

ACS Chemistry FROG

19

Chemical Energetics: Enthalpy

Chapter 7

Section 7.5. System and Surroundings


Learning Objectives for Section 7.5
Identify the system and the relevant surroundings for a given change.
Define and identify open, closed, and isolated systems.
Consider This 7.18. How do systems and surroundings interact?
Goal:
For a variety of examples, differentiate between the system and surroundings and determine
whether the systems are open, closed, or isolated.
Classroom options:
Allow a couple of minutes for students, working in small groups, to answer each part. After
each part, the groups can share their conclusions with the class and at the end of the activity
summarize the answers on the chalkboard or an overhead transparency.
This activity could be conducted as an open class discussion.
This activity could also be assigned as a homework problem and then discussed at the next
class section.
Time for activity:
Approximately 15-20 minutes, depending on the amount of discussion about interpretations of
systems and surroundings.
Instructor notes:
Show or refer students to Figure 7.6 and Table 7.1 as you begin the discussion of these
examples, so they are reminded of the definitions of system and surroundings and the
characteristics of isolated, closed, and open systems.
Help students see that there are different ways to look at these examples.
Students should reason and conclude:
(a) The relevant surroundings are the beaker, your hand, and the air surrounding the beaker.
Your hand feels cold because energy is being transferred from your hand to the beaker (as
well as from the warmer air to the cooler beaker) and the beaker is, in turn transferring
energy to the water as the NH4Cl(s) crystals dissolve. The system is closed because no matter
is exchanged with the surroundings.
(b) The relevant surroundings are the walls of the well-insulated container. The walls of the
container can transfer thermal energy to the system and become cooler. The system is closed.
If we assume that the walls of the container cannot transfer energy to or accept energy from
the system, then the system is isolated, because no exchange of matter, heat, or work occurs.
This is the assumption we make in using a simple calorimeter, like the one shown in Figure
7.7, where we assume that the expanded foam container neither gains nor gives up thermal
energy to its contents. This assumption cannot be strictly correct, but it is often good enough
to obtain results good to a few percent.
(c) For this case, we will assume that no energy transfers take place to or from the wellinsulated container. The water in the container is the surroundings for this example. The hot
piece of metal transfers thermal energy to the water. This transfer will continue until the
system (metal) and surroundings (water) reach the same temperature. When their
temperatures are the same, no net transfer of energy will occur between them. The system is
a closed system because only thermal energy is exchanged. No matter is exchanged.

20

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

(d) If the water and piece of metal together are the system (and no thermal energy is
transferred to or from the container walls), this is an isolated system, because no matter, heat,
or work enters or leaves this system. The surroundings (rest of the world) do not interact with
the system. The hot piece of metal will still transfer thermal energy to the water until the
metal and water reach the same temperature. When their temperatures are the same, no net
transfer of energy will occur between them. Since no thermal energy has entered or left this
system, the sum of the thermal changes in the system (cooling of the metal and warming of
the water) has to be zero. [This is the way we will analyze our calorimetric measurements.]
(e) The picture here shows figure in the Web Companion, Chapter 7, Section 7.6, page 1,
after the labels have been moved to differentiate the system from the surroundings for this
calorimeter set-up. This figure is very similar to what is shown in Figure 7.7 in the textbook,
with the exception that no stirrer is shown here and a syringe is not used to add a reagent in
the textbook figure.

[Note that the differentiation between system and surroundings in this figure is like the
differentiation in part (c) above. Alternatively, the solvent water could be taken to be part of
the system, as in part (d) above. That is the way we will treat the calorimeter in Section 7.6.]
Follow-up discussion:
Discuss important points in the "Reflection and Projection and use the discussion of part (e)
to lead into Section 7.6 on calorimetry.
Follow-up activities:
End of chapter problems 7.17 through 7.21.

ACS Chemistry FROG

21

Chemical Energetics: Enthalpy

Chapter 7

Section 7.6. Calorimetry and Introduction to Enthalpy


Learning Objectives for Section 7.6
Use the data from calorimetric measurements to calculate the thermal energy transferred to or
from a reaction system.
Use the data from constant pressure calorimetric measurements to calculate the enthalpy
change for the reacting system.
Investigate This 7.19. What happens when an acid and base are mixed?
Goal:
Observe that thermal energy is transferred from the reaction system to the surroundings when
an acid and a base are mixed.
Set-up time:
5-10 minutes, if the solutions are already available.
Time for activity:
Less than 5 minutes.
Equipment:
Small test tubes.
Long-stem plastic pipets.
Chemicals:
1 M NaOH (4.0 g per 100 mL of water). Dispense in properly labeled plastic pipets.
1 M HCl (8.5 mL concentrated HCl in 100 mL solution). Dispense in properly labeled plastic
pipets.
Procedure:
If your classroom situation is not safe for distributing the reagents to groups, ask for a few
volunteers to do the activity and report the results to the rest of the class.
SAFETY NOTE
Wear safety goggles.
Have each group/student add about 1 mL of the NaOH solution to the test tube and then about
1 mL of the HCl solution.
Gently touch the test tube near the bottom and report/record observations.
Alternative procedures:
Place 0.5 mL of NaOH on a liquid crystal thermometer at the position of the green stripe. Add
0.5 mL of HCl directly on the puddle of NaOH solution. Photographs of the results are:

base on thermometer
22

acid + base on thermometer


ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

NOTE: The color stripe on a liquid crystal thermometer is green when the temperature of the
surroundings matches the temperature printed on that stripe. If the temperature is more than
about 3-5 C hotter or colder than the temperature printed on a stripe, the stripe looks black
(because it is clear and the backing color shows through). If the stripe senses a temperature that
is just a little warmer than the printed temperature, it appears blue. If the stripe senses a
temperature that is just a little cooler than printed temperature, it appears red. In the example
shown in the photographs, the base appears to have a temperature of about toward the cooler
end of the 24-27 C range. When the acid is added the temperature of the liquid increases to
about the middle of the 27-29 C range. Thus, there is an increase in temperature of about 4 C.
A thermometer, liquid in glass or digital, could also be used to semi-quantify the temperature
change from the heat of reaction.
An extension or alternative activity might be to measure the temperature change when
ammonium nitrate or calcium chloride is added to water in a Styrofoam cup, thus linking
back to Investigate This 2.22 in Chapter 2, Section 2.5. The measured temperature changes
could then be used for the calculations in this section, in place of the ones in the textbook. It is
not a good idea to replace Investigate This 7.22, if you intend to do the extension of the ureadissolution activity in Chapter 9, Section 9.7.
Anticipated results:
The reaction between the NaOH and HCl solutions is exothermic, so the test tube should feel
warm (as we also see in the liquid crystal thermometer result above).
Follow-up discussion:
Use Consider This 7.20 to initiate discussion of this activity and Check This 7.21 as an
assignment to check understanding.
Follow-up activities:
Investigate This 7.22. What happens when urea dissolves in water?
End of chapter problems 7.22 through 7.27.
Consider This 7.20. What causes the changes when an acid and base are mixed?
Goal:
Conclude that the familiar (form Chapter 6) reaction of an acid and a base releases thermal
energy to the surroundings.
Classroom options:
This activity can be conducted as an open class discussion as Investigate This 7.19 is being
carried out.
Time for activity:
Approximately 5-10 minutes.
Instructor notes:
Students should reason and conclude:
The test tube felt warm after the acid was mixed into the base. Thermal energy is transferred
from the contents of the test tube to the surroundings, including fingers touching the test
tube. We recall from Chapter 6 that hydrochloric acid solution contains hydronium and
chloride ions and sodium hydroxide solution contains sodium and hydroxide ions. We know
that hydronium ions and hydroxide ions combine with one another to form water, so it seems
reasonable to conclude that this combination is the reaction that occurs when the solutions
are mixed and our results suggest that energy is released by the reaction. The energy

ACS Chemistry FROG

23

Chemical Energetics: Enthalpy

Chapter 7

increases the temperature of the solution and we detect the increased temperature because
thermal energy is transferred to our skin.
Follow-up discussion:
Discussion should help students reinforce or gain understanding of the difference between
systems and surroundings, thermal energy transfer, and the definition of enthalpy.
This is a good opportunity to review endothermic and exothermic changes.
Follow-up activities:
Check This 7.21. Is the reaction of an acid with a base exothermic or endothermic?
Investigate This 7.22. What happens when urea dissolves in water?
End of chapter problems 7.22 through 7.27.
Investigate This 7.22. What happens when urea dissolves is water?
Goal:
Determine the temperature change (decrease) when urea is dissolved in water.
Set-up time:
5-10 minutes.
Time for activity:
5-10 minutes.
Equipment:
Constant pressure calorimeter, shown in Figure 7.7. Good results can be obtained without a lid
or stirrer using the thermometer or a computer-interfaced temperature probe to stir the
contents of the Styrofoam cup.
100-mL graduated cylinder to measure the water.
Chemicals:
6.0 g of urea, H2NC(O)NH2(s).
Procedure:
Do this as a class investigation with volunteers to carry out the activity and have students
work in small groups to discuss and analyze the results.
SAFETY NOTE
Wear your safety goggles
Set up a simple constant-pressure calorimeter, like the one illustrated in Figure 7.7 (or the
simpler set-up suggested above), containing 100. mL of room temperature water.
While gently stirring the water, record its temperature every 15 seconds for 1-2 minutes.
Add 6.0 g of urea all at once to the water and stir vigorously to dissolve the urea as quickly as
possible, while continuing to record the temperature of the solution for another 3 minutes.
Anticipated results:
In numerous trials of this activity, using a computer interfaced temperature probe, the
temperature drops 3.5-3.6 C.
Follow-up discussion:
Use Consider This 7.23 to initiate discussion of the results of this activity as it is being carried
out.
Follow-up activities:
Consider This 7.24. What is observed for an endothermic reaction in a calorimeter?
Check This 7.25. Energy transfers in a calorimeter.
Worked Example 7.26. Determination of qP(reaction) for a reaction in a calorimeter.
Check This 7.27. Determination of qP(reaction) for dissolution of urea in a calorimeter.
24

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

Worked Example 7.28. Determination of Hreaction for a reaction in a calorimeter.


Check This 7.29. Determination of Hreaction for dissolution of urea in a calorimeter.
End of chapter problems 7.22 through 7.27.
Consider This 7.23. What is T when urea dissolves in water?
Goal:
Determine T for the solution when urea dissolves in water in a Styrofoam cup constantpressure calorimeter.
Classroom options:
Conduct this activity as an open class discussion as Investigate This 7.22 is done.
Time for activity:
Approximately 2 minutes in addition to Investigate This 7.22 for the class to agree on the
value for the temperature change.
Instructor notes:
Use the agreed upon results for Investigate This 7.22 to determine T.
Students should reason and conclude:
T = Tf TI = 3.5 to 3.6 C.
Follow-up activities:
Consider This 7.24. What is observed for an endothermic reaction in a calorimeter?
Check This 7.25. Energy transfers in a calorimeter.
Worked Example 7.26. Determination of qP(reaction) for a reaction in a calorimeter.
Check This 7.27. Determination of qP(reaction) for dissolution of urea in a calorimeter.
Worked Example 7.28. Determination of Hreaction for a reaction in a calorimeter.
Check This 7.29. Determination of Hreaction for dissolution of urea in a calorimeter.
End of chapter problems 7.22 through 7.27.
Consider This 7.24. What is observed for an endothermic reaction in a calorimeter?
Goal:
Treating the contents of an ideal calorimeter as an isolated system, give the signs for qP(reaction)
and qP(water) for an endothermic reaction in the calorimeter.
Classroom options:
Allow 2-3 minutes for students, working in small groups, to answer these questions. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
This activity could be conducted as an open class discussion.
Time for activity:
Approximately 10 minutes.
Instructor notes:
Introduce the constant-pressure calorimeter, using Figure 7.7 and/or the figure in the Web
Companion, Chapter 7, Section 7.6, page 1, as an example.
Discuss the concept of an ideal constant-pressure calorimeter and introduce the relationship
that is at the heart of our approach, equation (7.3): qP(reaction) + qP(solution) = 0.
NOTE: You (and some of your students) may be more familiar with a different approach to
ideal adiabatic calorimetry in which the reaction is taken as the system and the solution in which
it occurs is the surroundings. The problem that often arises for students in this approach is what
signs to assign to the thermal energy transfers (heats) from or to the system and surroundings. In

ACS Chemistry FROG

25

Chemical Energetics: Enthalpy

Chapter 7

the approach we have taken in this text, the calorimeter contents are regarded as an isolated
system, which transfers no energy to nor receives energy from the surroundings (including in
this ideal case, the calorimeter components themselves). Thus, the sum of all energy transfers
within the calorimeter must be zero. The observable variable in this system is the temperature of
the solution. The direction of the temperature change immediately provides the sign of
qP(solution), and, therefore, from equation (7.3), the sign of qP(reaction). This approach seems
effective in reducing amount of sign confusion in calorimetric calculations.
Students should reason and conclude:
(a) An endothermic reaction is one in which thermal energy is transferred to the reaction, so
qP(reaction) is positive.
(b) Since qP(reaction) is positive and qP(reaction) + qP(solution) = 0, qP(solution) must be negative, that is
thermal energy will leave the solution. This means that the solution will become cooler.
(c) The dissolution of urea in water is endothermic. The decrease in temperature observed in
Investigate This 7.22 is the evidence that the dissolution of urea is endothermic.
Follow-up discussion:
Review specific heat (Chapter 1, Worked Example 1.57, page 55) with respect to linking
thermal energy change to temperature change.
Follow-up activities:
Check This 7.25. Energy transfers in a calorimeter.
Worked Example 7.26. Determination of qP(reaction) for a reaction in a calorimeter.
Check This 7.27. Determination of qP(reaction) for dissolution of urea in a calorimeter.
Worked Example 7.28. Determination of Hreaction for a reaction in a calorimeter.
Check This 7.29. Determination of Hreaction for dissolution of urea in a calorimeter.
Consider This 7.30. How can you get more accurate values for Hreaction?
Check This 7.31. Effect of increasing the amount of another calorimetric reaction.
End of chapter problems 7.22 through 7.27.
Consider This 7.30. How can you get more accurate values for Hreaction?
Goal:
Students conclude that doubling the amount of reaction coupled with doubling the heat capacity
of the calorimeter contents gives the same temperature change and same precision for H(reaction)
as in the original trial.
Time for activity:
10-15 minutes.
Classroom options:
Allow 2-3 minutes for students, working in small groups, to answer these questions. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
This activity could be conducted as an open class discussion.
This activity could also be assigned as a homework problem and then discussed at the next
class section.
Instructor notes:
All the reasoning in this activity is based upon experimental values, not on any table values.
It is likely for students to reason that doubling the amount of reaction will double the energy
released by the reaction and hence should double the temperature change of the solution. The

26

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

experimental evidence contradicting this conclusion for this reaction system is designed to
produce the sort of cognitive dissonance that will put students in a receptive mode to learn
why their reasoning was wrong.
Students should reason and conclude:
(a) As indicated above, the result (same temperature change with double the amount of
reactants) is likely not to be what is expected.
(b) Doubling the amount of each reactant volume doubles both the energy released and the
heat capacity of the system. This leads to:
qP(reaction) = (m)(c)(T) = (400. g)(4.18 Jg1C1)(0.66 C1) = 1.10 103 J
qP(reaction) = qP(solution) = 1.10 103 J
The acetic acid, (0.200 L)(0.105 M) = 0.0210 mol, is still the limiting reactant
3
1.10 10 J
Hreaction =
= 52 kJmol1
0.0210 mol
This result is given to two significant figures because the energy value has an uncertainty of
about 2% (from the uncertainty in T), so Hreaction will have an uncertainty of about 2%.
This result (within the uncertainty) is the same as in Worked Example 7.28. There has been
no improvement in the precision of the result, because the temperature change is the same
and it is the limiting factor on the uncertainty in Hreactio.
(c) The example and questions in the Web Companion Chapter 7, Section 7.6, page 3, are
directly related to parts (a) and (b). Scaling the calorimeter up by a factor of three, as
suggested in the Companion, would result in the same temperature change although the
thermal energy transferred would increase by three. Thus, Hreaction would remain the same,
as it does in our example of a doubling of the scale of the experiment.
Follow-up discussion:
Discuss how to improve calorimetric measurements, such as making more precise temperature
measurements or changing the amount of reaction without changing the heat capacity, as
exemplified in Check This 7.31.
Follow-up activities:
Check This 7.31. Effect of increasing the amount of another calorimetric reaction.
End of chapter problems 7.22 through 7.27.

Section 7.7. Bond Enthalpies


Learning Objectives for Section 7.7
Define and give examples of homolytic bond cleavage reactions.
Write equations for homolytic bond cleavage and homolytic bond formation for compounds,
use bond enthalpies to calculate the enthalpy changes associated with these processes, and
obtain the enthalpy change for gas phase reactions.
Draw enthalpy level diagrams that show how bond enthalpies combine to give the enthalpy
change for a reaction.
Use bond enthalpies to predict whether, for given atoms, reactions will favor singly-bonded or
multiply-bonded products.

ACS Chemistry FROG

27

Chemical Energetics: Enthalpy

Chapter 7

Investigate This 7.32. What happens when yeast is added to hydrogen peroxide?
Goal:
Conclude that a gas (or gases) is produced in the exothermic (graduated cylinder gets warm)
reaction when yeast is added to hydrogen peroxide.
Set-up time:
10 minutes if all equipment and chemicals are already prepared.
Time for activity:
Approximately 10 minutes.
Equipment:
100 mL graduated cylinder.
Chemicals:
10 mL 3% hydrogen peroxide solution, H2O2.
1-2 mL liquid dishwashing detergent.
Approximately 2.0 g dried yeast. Instead of yeast, a small piece of liver or other fresh red
meat can be used.
Procedure:
This activity should be done by the instructor.
SAFETY NOTE
Wear your safety goggles.
Add hydrogen peroxide to the graduated cylinder followed by dishwashing detergent. Gently
swirl contents without creating suds.
Carefully add yeast to the solution. Try not to spill yeast granules on the sides of the cylinder.
Gently swirl the contents.
Feel the cylinder to determine if the cylinder feels warm, cool, or has not changed in
temperature.
Anticipated results:
Gas is produced which gradually fills the cylinder with soap-bubble foam. The detergent is
added so the bubbles will remain while Consider This 7.33 and Check This 7.34 are
completed before conducting Investigate This 7.35 to identify the gas.
The cylinder gets warm at the bottom where the reaction occurs, indicating that the reaction is
exothermic.
This photograph shows that the cylinder essentially fills with gas (foam) from the reaction.

28

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

Follow-up discussion:
Use Consider This 7.33 to initiate discussion about what reaction occurred and how to test for
the products of the reaction.
Students need to understand that chemical reactions involve breaking and making bonds.
Thus, energy is involved. In this example, the decomposition of hydrogen peroxide is slightly
exothermic. This means that the weaker bonds of the reactants are replaced by the stronger
bonds of the products.
Figure 7.8 (Bonding rearrangements for possible reactions of hydrogen peroxide) can be
displayed to discuss the possible reaction scenarios for the decomposition of hydrogen
peroxide.
Discuss how we can test for the possible products, leading students to observe Investigate
This 7.35.
Follow-up activities:
Check This 7.34. Properties of the possible H2O2 reaction products.
Investigate This 7.35. What is the gas formed in the hydrogen peroxide reaction?
Consider This 7.36. Which bonds break and form in the hydrogen peroxide reaction?
Consider This 7.33. What reaction occurs when yeast is added to hydrogen peroxide?
Goal:
Provide reasonable suggestions, based on the observations from Investigate This 7.32, about the
possible products of the reaction and ways to test these hypotheses.
Classroom options:
Allow 2-3 minutes for students, working in small groups, to answer these questions. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
This activity could be conducted as an open class discussion.
Time for activity:
Approximately 10-20 minutes, depending on how much of Check This 7.34 you wish to
include in the discussion.
Instructor notes:
You may need to introduce this discussion by pointing out that hydrogen peroxide is produced
in most aerobic organisms and can harm the organism, if it is not decomposed quickly.
Therefore, organisms, including yeast, contain enzymes (proteins) specifically designed to
decompose hydrogen peroxide. If you wish to demonstrate the presence of these enzymes in
other organisms, try small pieces of horseradish root, a higher organism than yeast. [Pieces of
red meat contain hemoglobin from the blood and the iron in this protein (see Figure 6.10) also
catalyzes the peroxide decomposition, as we will further explore in Investigate This 11.6,
Chapter 11, Section 11.1.] The bottom line here is that the product(s) are formed solely from
hydrogen peroxide as it decomposes in the presence of the yeast enzymes.
Students should reason and conclude:
We know a reaction occurred when yeast was added to hydrogen peroxide because bubbles of
gas formed and the graduated cylinder felt slightly warm to the touch.
Since only H and O atoms are available from H2O2, reasonable suppositions about the gas or
gases produced are H2 and/or O2. Molecular level models of formation of these gases are
shown in Figure 7.8 and symbolic equations for the reactions are given in the figure and in
equations (7.6) and (7.7).

ACS Chemistry FROG

29

Chemical Energetics: Enthalpy

Chapter 7

Students may know from previous study or can be prompted to figure out how to test for these
products, knowing that hydrogen burns in air (or explodes if it is mixed with air or oxygen)
and oxygen supports combustion. Figuring this out is the gist of Check This 7.34 and
Investigate This 7.35 involves carrying out the test.
Follow-up discussion:
Show or refer to Figure 7.8 to illustrate how the bonds might be broken and formed in the
decomposition reaction of hydrogen peroxide.
Follow-up activities:
Check This 7.34. Properties of the possible H2O2 reaction products.
Investigate This 7.35. What is the gas formed in the hydrogen peroxide reaction?
Investigate This 7.35. What is the gas formed in the hydrogen peroxide reaction?
Goal:
Students conclude, based on the results of a glowing splint test, that oxygen gas is the product of
the hydrogen peroxide decomposition reaction.
Set-up time:
A couple of minutes to arrange the safety shields for the test.
Time for activity:
3-5 minutes, including discussion of the outcome of the test.
Materials:
Foam in graduated cylinder from Investigate This 7.32.
Long fireplace match or wooden splint.
Stirring rod.
Procedure:
After discussing the background for testing for hydrogen gas or oxygen gas (covered in Check
This 7.34), test the flammability of the gas trapped in the foam from Investigate This 7.32.
This test should be carried out by the instructor following all safety precautions.
SAFETY NOTES
Wear your safety goggles.
Place the graduated cylinder behind a safety shield that
protects others as well as yourself when the gas is tested.
Very gently stir the foam to break up some of the bubbles, while keeping the gas in the
cylinder.
Light a long fireplace match or wooden splint.
Blow it out and quickly insert the glowing tip a few centimeters into the mouth of the
cylinder.
Observe and record the results.
Anticipated results:
The product is oxygen gas alone. Depending on how the glowing match is inserted in the
graduated cylinder, it might glow more brightly or reignite in the oxygen atmosphere, but it
could be wet from the soap bubbles and simply die out, which is why you try to get rid of as
much soap film as possible before the test. In any case, there is no explosion, so no
hydrogen/oxygen mixture has been produced. Here is a photograph of the result (with no
safety shields, because it was done with knowledge of the outcome and not in a classroom).

30

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

Clean-up:
After confirming that it is completely extinguished, the match can be thrown away. Rinse the
graduated cylinder with copious amounts of water.
Follow-up discussion:
Based upon the results of this activity, the correct reaction for the decomposition of hydrogen
peroxide is shown in Figure 7.8(a) and equation (7.6).
The glowing match glows more brightly or bursts into flame, indicating the presence of
oxygen gas. There is no "pop" (small explosion), so hydrogen gas is not present in the gas.
Use Consider This 7.36 to initiate discussion of which bonds are broken and which bonds are
formed in the reaction.
Follow-up activities:
Check This 7.37. Overall homolytic bond cleavage reaction of methane.
End of chapter problems 7.28 through 7.37.
Consider This 7.36. Which bonds break and form in the hydrogen peroxide reaction?
Goal:
Find, by modeling reactions (7.6) and (7.7), that the same number of two-electron bonds is
formed as is broken in each reaction.
Classroom options:
The activity could be conducted as an open class discussion with students building and
rearranging their models individually.
This activity could also be assigned as a homework problem and then discussed at the next
class section.
Time for activity:
10-15 minutes.
Instructor notes:
Show or refer students to reactions (7.6) and (7.7) as the activity is conducted.
Make classroom-size models of the reactants and products before the class session, as a
possible aid in the discussion.
Students should reason and conclude:
(a) For the reactants in reaction (7.6), three bonds must be broken: the two HO bonds in one
of the H2O2 molecules and the OO bond in the other. For the products, three bonds are
formed: two HO bonds (one each between a free H and the O in an OH to form H2O) and

ACS Chemistry FROG

31

Chemical Energetics: Enthalpy

Chapter 7

one OO bond to make the second bond between the OO left after the OH bonds were
broken. There are six two-electron bonds (represented by the plastic connectors) in both the
reactants and the products.
(b) For the reactants in reaction (7.7), two bonds are broken: both HO bonds. For the
products, two bonds are formed: one bond between the free Hs to form H2 and one OO bond
to make the second bond between the OO left after the OH bonds were broken.. There are
three two-electron bonds (represented by the plastic connectors) in both the reactants and the
products.
Follow-up discussion:
Use the discussion to introduce bond enthalpy and homolytic bond cleavage.
Follow-up activities:
Check This 7.37. Overall homolytic bond cleavage reaction of methane.
Consider This 7.38. How do bond enthalpies for single and multiple bonds compare?
Check This 7.39. Formation of polyethylene.
Check This 7.40. Sulfur-sulfur bonds.
Worked Example 7.41. Reaction enthalpy from average bond enthalpies.
Check This 7.42. Reaction enthalpy from average bond enthalpies.
Check This 7.43. The hydrogen peroxide decomposition reaction.
End of chapter problems 7.28 through 7.37.
Consider This 7.38. How do bond enthalpies for single and multiple bonds compare?
Goal:
Conclude that there are two patterns of relative strengths of multiple compared to single bonds.
Classroom options:
Allow 2-3 minutes for students, working in small groups, to answer these questions. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
This activity could be conducted as an open class discussion.
Time for activity:
Approximately 10 minutes.
Instructor notes:
Remind students that only elements for which multiple bond data are given in Table 7.3 (that
is, CC, NN, OO, and SS) can be analyzed for this activity.
Table 7.4 and Figure 7.9 display the patterns to be found in this activity. Show or refer
students to Figure 7.9 after they have discussed their results and see how the pattern(s) they
found compare.
Students should reason and conclude:
(a) There are two patterns of relative strengths of multiple bonds compare to single bonds in
the data in Table 7.3. For carbon and sulfur, the average bond enthalpies of multiple bonds
are less than the sum of the average bond enthalpies of single bonds with an equivalent
number of bonding electrons. For example, a S=S bond with four bonding electrons has an
average bond enthalpy of only 352 kJmol1, but two SS bonds with a total of four bonding
electrons have a total bond enthalpy of 2 268 kJmol1 = 536 kJmol1. The double bond is
much weaker than two single bonds. For nitrogen and oxygen, the opposite pattern is
observed, the average bond enthalpies of multiple bonds are greater than the sum of the
average bond enthalpies of single bonds with an equivalent number of bonding electrons.

32

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

(b) Two patterns emerge, as indicated in part (a). These patterns are given quantitatively for
carbon, nitrogen, and oxygen in Table 7.4 and illustrated graphically in Figure 7.9.
Follow-up activities:
Check This 7.39. Formation of polyethylene.
Check This 7.40. Sulfur-sulfur bonds.
Worked Example 7.41. Reaction enthalpy from average bond enthalpies.
Check This 7.42. Reaction enthalpy from average bond enthalpies.
End of chapter problems 7.28 through 7.37.

Section 7.8. Standard Enthalpies of Formation


Learning Objectives for Section 7.8
Define standard states for elements and compounds and write the equations whose enthalpy
changes are the standard enthalpies of formation of the compounds.
Use standard enthalpies of formation to calculate the standard enthalpy change for a reaction.
Draw enthalpy level diagrams that show how standard enthalpies of formation combine to
give the standard enthalpy change for a reaction.

Section 7.9. Harnessing Energy in Living Systems


Learning Objectives for Section 7.9
Define coupled reactions and identify examples based on the definition.
Show whether coupling between two reactions would be an energetically favorable
combination.

Section 7.10. Pressure-Volume Work, Internal Energy and Enthalpy


Learning Objectives for Section 7.10
Identify the forms of energy transferred in physical and chemical changes and show how
energy is conserved in the changes.
Draw molecular level representations of thermal energy (undirected kinetic energy) and
mechanical energy (directed kinetic energy) transfers.
Define and identify the variables that are functions of state and those that are functions of the
path for a given change.
State the first law of thermodynamics in terms of internal energy, heat, and work and use it to
analyze a change that occurs by different pathways.
Use pressure and volume data or the ideal gas equation to calculate the pressure-volume work
done in a reaction that involves gases.
Consider This 7.55 How is the first law of thermodynamics used?
Goal:
Use the first law of thermodynamics to analyze the operation of a small engine.
Classroom options:
Allow 2-3 minutes for students, working in small groups, to work on this activity. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
This activity could be conducted as an open class discussion.
ACS Chemistry FROG

33

Chemical Energetics: Enthalpy

Chapter 7

This activity could also be assigned as a homework problem and then discussed at the next
class section.
Time for activity:
From 5 to 15 minutes, depending on the depth of the discussion of the first law and the
relationship of the state function, E, and the path dependent variables, q and w.
Instructor notes:
Students might need some hints for this activity. In particular, they need to be aware that E is
the same for the original and the redesigned engine because both engines are working between
the same initial and final states.
Students should reason and conclude:
For the original engine, heat and work are produced in the surroundings and we have q =
5 kJ (per power stroke) and w = 2 kJ (per power stroke). Thus E = q + w = 7 kJ (per
power stroke).
For the redesigned engine, more work is performed, w = 3 kJ (per power stroke) but E is
unchanged, because the redesigned engine operates between the same initial and final states
as the original engine and E is a function of state (dependent only on the initial and final
state of the system -- the engine). Therefore,
E = 7 kJ (per power stroke) = q + w
q = E w
q = [7 kJ (per power stroke)] [ 3 kJ (per power stroke)] = 4 kJ (per power stroke).
The same amount of energy is released in the redesigned engine. More work is done, but less
heat is released. Overall, the redesigned engine is more efficient (more of the energy is
released as work to do useful things in the surroundings).
Follow-up discussion:
Students should be able to explain why the redesigned engine is more efficient and why E is
the same for both engines.
Follow-up activities:
Worked Example 7.56. Pressure-volume work at constant pressure.
Check This 7.57. Pressure-volume at constant pressure.
Consider This 7.58. How can E (or H) be the same for a reaction at constant volume and
constant pressure
End of Chapter problems 7.50 through 7.58.
Consider This 7.58. How can E (or H) be the same for a reaction at constant volume
and constant pressure?
Goal:
Students show how q and w combine to give the same E (or H) for constant volume and
constant pressure reactions.
Classroom options:
Allow 3-5 minutes for students, working in small groups, to work on this activity. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
This activity could be conducted as an open class discussion.
Instructor notes:
As you introduce this activity, remind students of the sign conventions for q and w.
Students should reason and conclude:
34

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

(a) The reaction is exothermic, that is, thermal energy leaves the reacting system, so thermal
energy transfer to the system is negative (< 0).
(b) More thermal energy enters the solution when the reaction is run at constant volume (the
temperature change in the solution is larger) than when run at constant pressure. Thus, more
thermal energy leaves the reacting system at constant volume than at constant pressure. We
can write this relationship as |qV| > |qP|. Or we can say that, if qP < 0, then qV << 0, that is, qV
is more negative than qP or, conversely, qP is more positive than qV.
(c) Since a gas is produced by the reaction, the change in volume at constant pressure is
positive, V > 0. Thus, the work done on the system, w = PV, is negative, w < 0.
(d) Under constant volume conditions, E = qV. Under constant pressure conditions,
E = qP + w. If E is the same for both reaction conditions, then we must have qV = qP + w.
If we substitute the numerical values in this equation, we get: (<< 0) = (< 0) + (< 0). This
equation can be satisfied because two negative quantities, qP and w, combine additively to
give a larger negative quantity. Thus, E can be (and is) the same for both reaction
conditions. The same sort of analysis is possible for H, which can be written H = qP or
H = E w = qV w. You can show how H can be a constant.
Follow-up activities:
End of Chapter problems 7.57 through 7.63.

Section 7.11. What Enthalpy Doesn't Tell Us


Learning Objectives for Section 7.11
Determine whether a process is consistent with (allowed) by the first law of thermodynamics.

Section 7.13. Extension: Ideal Gases and Thermodynamics


Learning Objectives for Section 7.13
Use the kinetic-molecular model of gases to explain the observed effects of changes in P, V, n,
or T for gases.
Calculate the final value of P, V, n, or T for an ideal gas, given their initial values and the
changes in three of the variables.
Use the ideal gas equation to calculate the P-V work done on or by a chemical reaction
system.
Investigate This 7.63. What happens when a gas is compressed or heated?
Goal:
Observe what happens when a sample of gas is compressed or heated.
Set-up time:
15 minutes.
Time for activity:
10-15 minutes.
Materials:
50-mL plastic syringe.
U-tube made of plastic tubing (see set-up in text). A completely adequate U-tube manometer
is readily made from about 1.5 m of clear plastic tubing such as Tygon. The tubing can be
clamped into a U shape or taped in a U shape to an appropriate length of 1-by-4 lumber.
ACS Chemistry FROG

35

Chemical Energetics: Enthalpy

Chapter 7

About 20-30 cm of one end of the tubing should not be part of the U, but left, as in the sketch,
for attaching the sample syringe.
NOTE: It is, of course, possible to use pressure transducers interfaced to a computer to do this
investigation and project the pressures. However, the action of the gas in pushing down the
liquid in the manometer is then lost and the impact of the investigation is probably diminished.
Chemicals:
Food coloring.
Water.
Procedure:
This is a class activity that can be done with student volunteers.
SAFET NOTE
Wear your safety goggles.
Draw enough colored water (red is usually best for visibility) into the tube to fill the U to
about the half-way point in each arm.
When the syringe is attached, the equal pressure in the arms will probably be upset, so you
have to manipulate the plunger to bring the pressure on the syringe side to atmospheric.
To avoid making a fountain, take care not to push the plunger in too rapidly or too far when
carrying out the investigation.
Push the syringe plunger in to decrease the volume of air in the syringe by several milliliters.
Record your observations on the liquid levels in the U-tube.
Readjust the syringe plunger until the U-tube liquid levels are the same again. While holding
the volume constant, warm the syringe by holding it in your hand or directing a warm stream
of air onto it. Record your observations on the liquid levels in the U-tube.
Clean-up:
Rinse the tube with water, so as not to permanently discolor it and make it harder to use again
for this activity.
Follow-up discussion:
Use Consider This 7.64 to initiate discussion, which can be begun while the activity is carried
out.
Follow-up activities:
Consider This 7.65. How are manometer readings interpreted and quantified?
Consider This 7.66. How does kinetic-molecular theory apply to Investigate this 7.63?
Check This 7.67. Relationships among the properties of a gas sample.
Worked Example 7.68. Using the ideal-gas equation.
Check This 7.69. Using the ideal-gas equation.
Worked Example 7.70. Using the ideal-gas equation.
Worked Example 7.71. Using the ideal-gas equation.
End of chapter problems 7.63 through 7.70.
Consider This 7.64. What causes changes when a gas is compressed or heated?
Goal:
Connect changes in fluid levels in the manometer in Investigate This 7.63 to changes in the
pressure of the gas sample trapped on one side of the manometer and suggest a molecular level
explanation for why compression or warming of the gas changes the pressure in the observed
direction.

36

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

Classroom options:
Allow 3-5 minutes for students, working in small groups, to work on this activity. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
This activity could be conducted as an open class discussion.
Instructor notes:
Be sure the class agrees on the observations from Investigate This 7.63 as they are discussed.
Students should reason and conclude:
(a) The liquid level in the sample (syringe) arm of the manometer moves down when plunger
is pushed in decreasing the volume of gas in the syringe. This must mean that more pressure
is being exerted on the surface of the liquid in this arm, compared to the pressure before the
gas volume was decreased. When the plunger is pushed in, the volume available to the gas
decreases, so there are more molecules per unit volume. The number of molecules colliding
with the walls, including the surface of the manometer liquid, is increased and, hence, the
pressure is increased.
(b) The liquid level in the sample (syringe) arm of the manometer moves down when the gas
is warmed while keeping its volume the same. This must mean that more pressure is being
exerted on the surface of the liquid in this arm, compared to the pressure before the gas was
warmed. When the gas in the syringe is warmed, the average kinetic energy of the molecules
increases and the average speed with which they collide with the walls, including the surface
of the manometer liquid is increased, and, hence, the pressure is increased. The liquid moves
down in the manometer arm.
Follow-up discussion:
Show or refer students to Figure 7.21 to discuss how a manometer works as this activity is
being carried out. This discussion should lead into Consider This 7.65 to see how manometer
readings are interpreted and quantified.
Follow-up activities:
Consider This 7.65. How are manometer readings interpreted and quantified?
Consider This 7.66. How does kinetic-molecular theory apply to Investigate this 7.63?
Check This 7.67. Relationships among the properties of a gas sample.
Worked Example 7.68. Using the ideal-gas equation.
Check This 7.69. Using the ideal-gas equation.
Worked Example 7.70. Using the ideal-gas equation.
Worked Example 7.71. Using the ideal-gas equation.
End of chapter problems 7.63 through 7.70.
Consider This 7.65. How are manometer readings interpreted and quantified?
Goal:
Calculate the pressure represented by the reading of a water manometer.
Classroom options:
Allow 3-5 minutes for students, working in small groups, to work on this activity. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
This activity could be conducted as an open class discussion.

ACS Chemistry FROG

37

Chemical Energetics: Enthalpy

Chapter 7

Instructor notes:
Show or refer students to the picture of the set up from Investigate This 7.63 to begin this
activity and then Figure 7.21 as they work on the activity.
Students should reason and conclude:
(a) The liquid level in the manometer is higher in the sample arm than in the arm open to the
atmosphere. This is the opposite situation to that shown in Figure 7.21 and must mean that
the pressure of the gas sample in the syringe is lower than atmospheric pressure. You must
have moved the syringe plunger out, increasing the volume of air in the syringe and thus
lowering its pressure. There are fewer molecules per unit volume and, therefore fewer
molecules colliding with the walls of the container and the surface of the fluid in the
manometer. Hence, the pressure is lower than before the plunger was moved.
(b) There is more than one strategy for finding the pressure of gas in the syringe. Here is one
approach. The quantity h [see Figure 7.21(b)] measures the level of liquid in one arm of the
manometer compared to the level in the other. Let us define h = Lopen Lsample. where Lopen is
the level in the arm open to the atmosphere and Lsample is the level in the arm attached to the
sample. When Lopen > Lsample, then h > 0, and the pressure of the sample given by the equation
in Figure 7.21(a) is greater than atmospheric, which we know is true under these conditions.
For the problem here, we are told that Lopen < Lsample, so h < 0, and the data in the problem
give h = 15.7 cm = 0.157 m. Thus, we have:
Psample = Patm + gh
1 kg
(9.80 ms2)(0.157 m) = 98.9 kPa
Psample = 100.4 kPa +
1 L

The calculated sample pressure is lower than atmospheric pressure, which we know is true
from our reasoning in part (a).
Follow-up discussion:
Use the discussion and reasoning for this problem as an opening to introduce the kineticmolecular model of gases.
Follow-up activities:
Consider This 7.66. How does kinetic-molecular theory apply to Investigate this 7.63?
Check This 7.67. Relationships among the properties of a gas sample.
Worked Example 7.68. Using the ideal-gas equation.
Check This 7.69. Using the ideal-gas equation.
Worked Example 7.70. Using the ideal-gas equation.
Worked Example 7.71. Using the ideal-gas equation.
End of chapter problems 7.63 through 7.70.
Consider This 7.66. How does kinetic-molecular theory apply to Investigate This 7.63?
Goal:
Use the kinetic-molecular theory to explain the results of Investigate This 7.63.
Classroom options:
Allow 3-5 minutes for students, working in small groups, to work on this activity. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
This activity could be conducted as an open class discussion.

38

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

Instructor notes:
Introduce and discuss the kinetic-molecular theory before conducting this activity.
The previous explanations in Consider This 7.64 and 7.65 have been given in terms of the
kinetic-molecular theory because some students will probably know something about the
behavior of gases from previous courses and you can capitalize on this by interweaving an
introduction to kinetic-molecular theory as these activities are discussed.
Students should reason and conclude:
(a) The number of molecules per unit volume increases as the volume of air in the syringe
decreases. The number of molecules hitting the walls per unit time increases, which we
observe as an increase in pressure of the gas.
(b) Warming the gas sample causes the average speed of molecules to increase, which means
that their collisions with the wall will be more energetic. We observe this increase in the
energy of the wall collisions as an increase in pressure of the gas.
(c) The pressure of the gas will decrease. Removing some of the molecules while keeping the
volume constant means that there will be fewer molecules per unit volume [as in part (a)
where we decreased the density of the gas by increasing its volume]. Fewer molecules will
collide with the walls of the container, thus decreasing the push on the walls and hence
decreasing the pressure.
Follow-up discussion:
Use the discussion of this activity as a lead-in to the ideal-gas equation.
Follow-up activities:
Check This 7.67. Relationships among the properties of a gas sample.
Worked Example 7.68. Using the ideal-gas equation.
Check This 7.69. Using the ideal-gas equation.
Worked Example 7.70. Using the ideal-gas equation.
Worked Example 7.71. Using the ideal-gas equation.
End of chapter problems 7.63 through 7.70.
Investigate This 7.72. Do reactions in open containers and capped containers differ?
Goal:
Observe that the temperature change of the solution in the reaction of hydrochloric acid with
sodium hydrogen carbonate (baking soda) is different when the reaction takes place in a closed
compared to an open container.
NOTE: The analysis of this system is relatively straightforward, but does demand keeping close
track of signs and directionality of effects. Since the bottom line message of the body of the
chapter is that E H for reactions in solution, you and your class might find it confusing that
effort is expended on analysis of a gas-producing system where finding the difference between
E and H, providing an application of the ideal gas equation, is the objective. If you think that
the pedagogical value of the analysis is worthwhile, the deviation of this system from the
mainline message should be discussed. Here, a gas is produced, so the volume change is
considerable, and the energy change is rather small. As a consequence the PV work term is
relatively important compared to the heat term and a difference between E and H, about
10%, is rather easy to observe. You can do the activity and analyze it only qualitatively
(directionality) as an assessment of student understanding of the directionality of the changes in
thermodynamic variables.

ACS Chemistry FROG

39

Chemical Energetics: Enthalpy

Chapter 7

Set-up time:
15-20 minutes. You should add the weighed sodium hydrogen carbonate (baking soda) to dry
plastic soda bottles before class. There is no point taking up class time for this part of the set
up. Two versions of the set up are given here. Dr. Jonathan Mitschele, Saint Josephs College,
Standish, ME, found that the set up described and pictured in the text is more elaborate than it
needs to be. Both the original and his alternative version are outlined.
NOTE: If neither of these options appeals to you, the animation in the Web Companion, Chapter
7, Section 7.11, is an idealized representation of this activity with data for further analysis.
Time for activity:
5-10 minutes. Its best to have two people available to carry out this activity.
Materials:
Two dry 2-L soft drink bottles. Smaller 1-L bottles would work just as well, if a thermometer
short enough to fit completely in them is available.
Cap for a soft drink bottle (for the alternative set up).
Two 10 to 110 C thermometers. One of these thermometers must fit completely into the soft
drink bottles. Shorter range thermometers are OK as well, if the range includes 0 to 25 C.
Watch or timer.
50-mL plastic syringe.
Needle for the syringe (for the original set up).
Short length of small diameter Tygon tubing to attach to the syringe outlet (for the
alternative set up).
New 24/40 ribbed rubber septum (for the original set up).
Length of 18-gauge copper wire to hold the septum on the bottle (for the original set up).
Powder funnel.
A 0.1-g (or better) balance.
Chemicals:
Two 8.4-g samples of sodium hydrogen carbonate [baking soda, NaHCO3(s)].
Two 20-mL samples of 6 M hydrochloric acid (HCl) solution.
This picture shows the materials and reagents for the original version of the constant volume
activity:

Original procedure:

SAFETY NOTE
Wear your safety goggles.

40

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

Before class, add 8.4 g of NaHCO3(s) [= NaHOCO2(s)] to each of the two dry 2-L bottles.
Use a powder funnel and take care not to get the solid on the walls of the bottle. It should all
be at the bottom of the bottles.
Place a thermometer in one of the bottles (with the thermometer bulb in one of the depressions
in the bottom of the bottle, if this is the style of bottle you have).
Use the other thermometer to measure the temperature of the 6 M HCl(aq) to the nearest
0.1 C. Record the temperature. (It is wise have two people carrying out this activity. While
one manipulates the 2-L bottle, the other can measure and record the temperature of the HCl
solution, draw 20 mL into the syringe, and add it to the reaction vessel.)
Use the syringe to dispense 20 mL of the HCl solution into the 2-L bottle.
Record the temperature and time as you swirl the mixture to be sure all of the solid contacts
the acid solution.
Tilt the bottle to bring all the solution together around the thermometer bulb and make sure it
is submerged in the solution.
Record the temperature of the mixture approximately every ten seconds for about two
minutes.
Repeat the experiment with the second sample with following changes:
After the NaHCO3(s) and thermometer are in the bottle, seal the bottle with the septum. Wire
the septum to the bottle. Although these septa fit rather tightly and the collar is stretched over
the bottle threads, a modest pressure builds up in the bottle and wiring the septum on helps
insure that there will be no leaks as well as that it will stay put. For best results, a new septum
should be used each time the closed reaction is conducted.
Put the syringe needle through the septum to inject the acid into the bottle. The picture in the
textbook shows the set up at this point (without the septum wired on, as it should have been).
IMPORTANT: When the reaction is complete, carefully remove the needle from the syringe.
Insert the needle through the septum to release the gas pressure from inside the bottle before
removing the septum.
Alternative procedure:
SAFETY NOTE
Wear your safety goggles.
Before class, add 8.4 g of NaHCO3(s) to each of the two dry 2-L bottles. Use a powder funnel
and take care not to get the solid on the walls of the bottle. It should all be at the bottom of the
bottles.
Place a thermometer in one of the bottles.
Use the other thermometer to measure the temperature of the 6 M HCl(aq) to the nearest
0.1 C. Record the temperature. (It is wise have two people carrying out this activity. While
one manipulates the 2-L bottle, the other can measure and record the temperature of the HCl
solution, draw 20 mL into the syringe, and add it to the reaction vessel.)
Carefully tilt the 2-L bottle on its side, keeping all the solid at the bottom, with the bottom
inclined a little above the top.
Use the syringe with attached short length of Tygon tubing to add 20 mL of HCl solution
near the top of the bottle, so it does not contact the solid. Note that it would not be necessary
to take such care to add the acid to this uncapped reaction, but it is done to be sure that the
conditions are essentially all the same for the uncapped and capped reactions.

ACS Chemistry FROG

41

Chemical Energetics: Enthalpy

Chapter 7

Turn the bottle upright, swirl to get all the solid and acid mixed, and read and record the
temperature of the solution as soon as possible.
Tilt the bottle to bring all the solution together around the thermometer bulb and make sure it
is submerged in the solution.
Record the temperature of the mixture approximately every ten seconds for about two
minutes.
Repeat the experiment with the second sample with following changes:
After the HCl solution has been added to the bottle, seal the bottle with the bottle cap while
the bottle is still inclined and the acid and solid have not come in contact. When the cap is
securely in place, continue the procedure as above.
When the readings have all been taken, carefully loosen the bottle cap a bit to release the gas
pressure in the bottle.
Clean-up:
Rinse the bottles and syringe with copious amounts of water.
Sample results:
In a sample set of reactions, the initial temperature of the 6 M HCl solution was 19.5 C
(temperatures read to the nearest 0.5 C). Temperature vs. time data, beginning 10 seconds
after addition of the HCl solution, are given in this graph.
10

Temperature, deg C

capped

uncapped

20

40
5

60
80
Time, sec

100

120

We see that there is a difference of 2.0 C between the capped and uncapped reactions. The
maximum changes in temperature, T for the capped and uncapped reactions are 11.5 C and
13.5 C, respectively.
NOTE: These results were obtained using an earlier version of the activity in which only 6.8 g
of NaHCO3(s), 0.081 mol, was reacted instead of the 0.10 mol in the present version.
Extrapolating to 0.10 mol of reaction, we should observe T for the capped and uncapped
reactions as 14.2 C and 16.7 C, respectively. These are the values we will use for further
calculations based on Investigate This 7.72. The values from the Web Companion are 15.5 C
and 18.8 C, respectively, for the capped and uncapped reactions.
Follow-up discussion:
Use Consider This 7.73 to initiate discussion of the results from this investigation. The
analysis is carried on through all the rest of the activities in this section, which are scaffolded
to lead in steps to the final result.
42

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

Follow-up activities:
Consider This 7.74. What are qV and qP for the hydrogen carbonate-acid reaction?
Worked Example 7.75. Converting qP (or H) to qV (or E).
Consider This 7.76. What is w for the hydrogen carbonate-acid reaction?
Check This 7.77. Ereaction and Hreaction for the hydrogen carbonate-acid reaction.
End of chapter problems 7.63 through 7.70.
Consider This 7.73. Why do reactions in open and capped containers give different
results?
Goal:
Identify the constant volume and constant pressure systems in Investigate This 7.72 and reason
that this difference causes a difference in the observed temperature changes.
Classroom options:
Allow 3-5 minutes for students, working in small groups, to work on this activity. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
Instructor notes:
Be sure that the class agrees on the results from Investigate This 7.72 as you begin this
discussion. You should use your own results, if you have them, or those from the reactions
reported above for Investigate This 7.72, or the data from the Web Companion, Chapter 7,
Section 7.11.
Students should reason and conclude:
(a) The maximum change in temperature in the open bottle was 16.7 C (or whatever the
class observed in their activity). The maximum change in temperature in the open bottle was
14.2 C (or whatever the class observed in their activity). The endothermic nature of this
reaction is a bit surprising, because other acid-base reactions observed or discussed earlier in
the chapter were exothermic. A difference in the hydrogen carbonate-acid reaction might be
that a gas is produced, as well as water, from the acid-base reaction.
(b) The reaction in the capped bottle was carried out at constant volume, the volume of the
bottle in which all products were trapped. The reaction in the open bottle was carried out at
constant pressure. In this latter case, the gas produced by the reaction could push back the air
in the bottle, which could leave the bottle, thus keeping the pressure in the bottle at
atmospheric pressure.
(c) In a constant volume reaction system, no work is done by or on the system, but in a
constant pressure system that produces a gas, work can be done by the system on its
surroundings by pushing them back. Perhaps it is this difference in the work produced that
causes the difference in the temperature changes in the two systems.
Follow-up discussion:
Try to be sure the discussion clearly makes these points: (1) the reaction in the capped bottle
is carried out at constant volume (no work is done) and (2) the reaction in the uncapped bottle
is carried out at constant pressure and the carbon dioxide does work by pushing out the air,
that is, pushing back the atmosphere.
Follow-up activities:
Consider This 7.74. What are qV and qP for the hydrogen carbonate-acid reaction?
Worked Example 7.75. Converting qP (or H) to qV (or E).
Consider This 7.76. What is w for the hydrogen carbonate-acid reaction?
ACS Chemistry FROG

43

Chemical Energetics: Enthalpy

Chapter 7

Check This 7.77. Ereaction and Hreaction for the hydrogen carbonate-acid reaction.
End of Chapter problems 7.63 through 7.70.
Consider This 7.74. What are qV and qP for the hydrogen carbonate-acid reaction?
Goal:
Calculate qV and qP for the hydrogen carbonate-acid reaction observed in Investigate This 7.72.
Classroom options:
Allow 3-5 minutes for students, working in small groups, to work on this activity. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.
This activity could also be assigned as a homework problem and then discussed at the next
class session.
Instructor notes:
If necessary, briefly review calorimetric calculations.
Students should reason and conclude:
In addition to the assumptions that the solution exchanges no thermal energy with the
surroundings during the reaction and that the solution has a specific heat = 4.18 J g-1 C-1, we
will assume that the density of the hydrochloric acid solution is 1.00 g mL-1. Thus, the mass
of solution is 20 g plus the mass of the added sodium hydrogen carbonate (8.3 g) minus the
mass of gas, carbon dioxide, lost (0.1 mol = 4.4 g), which is a total of 24 g. For our analysis
we know that qreaction + qsolution = 0, or rearranging, we have qreaction = qsolution. For the
constant volume system, qV = qreaction and for the constant pressure system, qP = qreaction,.
For the constant volume system, T = 14.2 C and we get
qV = qsolution = (m)(c)(T) = (24.0 g) (4.18 J C-1 g-1) (16.7 C) = 1.42 kJ
For the constant pressure system, T = 16.7 C and we get
qP = qsolution = (24.0 g) (4.18 J C-1 g-1) (16.7 C) = 1.68 kJ
[The difference between these two values, 0.26 kJ, represents the pressure-volume work done
by the system on the surroundings (as the gas is produced and pushes the atmosphere back).
The work done on the system is 0.26 kJ. This is energy lost by the system, which is why
more energy has to be supplied to the system from the thermal surroundings (the solution) for
the constant pressure reaction.]
Follow-up activities:
Worked Example 7.75. Converting qP (or H) to qV (or E).
Consider This 7.76. What is w for the hydrogen carbonate-acid reaction?
Check This 7.77. Ereaction and Hreaction for the hydrogen carbonate-acid reaction.
End of Chapter problems 7.63 through 7.70.
Consider This 7.76. What is w for the hydrogen carbonate-acid reaction?
Goal:
Use ideal gas behavior to calculate w for the hydrogen carbonate reaction and compare the
result with the difference between qV and qP from Consider This 7.74.
Classroom options:
Allow 3-5 minutes for students, working in small groups, to work on this activity. Then, the
groups can share their conclusions with the class, summarizing answers on the chalkboard or
an overhead transparency.

44

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

This activity could also be assigned as a homework problem and then discussed at the next
class session.
Instructor notes:
If necessary, review limiting reactant calculations as an introduction to this activity.
Students should reason and conclude:
(a) The molar mass of sodium hydrogen carbonate, NaHCO3, is 84 gmol1, so for our
activity we have
8.4 g
8.4 g NaHCO3 =
= 0.10 mol NaHCO3
84 g mol 1
(b) We used 20 mL of 6 M HCl (= 6 M H3O+), which gives
0.020 L of 6 M H3O+ = (0.20 L)(6 molL1) = 0.12 mol of H3O+
(c) Equation (7.57) shows that the hydronium ion and sodium hydrogen carbonate react in a
one-to-one molar ratio to produce one mole of carbon dioxide gas for each mole of each
reactant that reacts. The limiting reactant in this activity is the sodium hydrogen carbonate,
since there are fewer moles of this reactant. Thus, 0.10 mol of NaHCO3(s) produces 0.10 mol
of CO2(g).
(d) Using the ideal gas equation to determine PV for the production of 0.10 mol of gas, we
get
PV = nRT = (0.10 mol) (8.314 J mol-1 K-1) (293 K) = 0.24 kJ
(e) Substituting our result from part (d) into equation 7.58 gives
qV = qP PV = qP nRT = qP 0.24 kJ
Rearranging, we get
qP qV = 0.24 kJ
The value we got for this difference in Consider This 7.74 was 0.26 kJ. These results are
essentially the same, especially considering the crudeness of the experimental set-up in
Investigate This 7.72, and show that equation 7.58 is satisfied in this reaction system.
Follow-up activities:
End of Chapter problems 7.63 through 7.70.

ACS Chemistry FROG

45

Chemical Energetics: Enthalpy

Chapter 7

Solutions for Chapter 7 Check This Activities


Check This 7.6. The difference between warm and cool gases
The motion tails on the molecules represented in Figure 7.2 are supposed to suggest their
relative speeds. The tails on the gases inside and at the top of the chimney have longer tails than
those outside at the bottom. Thus, the molecules that have not reacted or entered the chimney
are moving more slowly and we have said that molecular motion and temperature are
proportional, so the slower moving molecules are cooler than those in the chimney or leaving it.
Check This 7.16. The system in Figure 7.5
The molecules in the skin are system. They are made to move faster by being pushed on as the
hands are rubbed briskly together or by absorbing infrared energy emitted by the flame.
Check This 7.17. Closed and open systems
(a) The fuel and flame are an open system, because mass, oxygen for the combustion, must be
supplied from the surroundings.
(b) Explosions like this are a closed system. In an explosion, thermal energy is released and
work (mechanical energy) is done on the surrounding (by pushing them back). However, no
mass has to be transferred to the system, because all of the reactant atoms are included in the
ammonium nitrate molecules. The products can, of course, mix with the surroundings, but this
is not relevant to the reaction and its consequences.
Check This 7.21. Is the reaction of an acid with a base exothermic or endothermic?
(a) The solutions used in Investigate This 7.19 contain ions formed by the reagents dissolved in
water: NaOH(aq) is a solution of Na+(aq) and OH(aq) and HCl(aq) is a solution of H3O+(aq)
and Cl(aq). The reactive species are H3O+(aq) and OH(aq) and the reaction is:
H3O+(aq) + OH-(aq) 2H2O(aq)
(b) The reaction is exothermic. When the reactant are mixed, the test tube becomes warm to the
touch, so thermal energy must be leaving the reacting system and warming the surroundings,
including fingers touching the test tube. The enthalpy of the system decreases, Hfinal < Hinitial and
H is negative.
Check This 7.25. Energy transfers in a calorimeter
The thermometer indicates that the temperature of the solution increases. Solvent water
molecules (and other molecules and ions dissolved in the water) next to reacting ions, have
absorbed energy by contact with the products of the chemical reaction, which have been
energized by the reaction. These energized water molecules come in contact with other water
molecules and transfer some of their energy to them and so on. Warmer water is less dense than
cooler water (above 4 C), so convection currents are set up in the liquid that mix the warmer
with the cooler water and enhance further contacts between molecules for energy exchange.
Eventually, water molecules that have gained energy come in contact with the thermometer.
Transfer of thermal energy via contact with the thermometer occurs, causing the mercury atoms
to vibrate faster, increasing its volume. The expanding volume of the mercury results in the
mercury filling more of the volume of its container (the narrow tube of the thermometer) and
indicating that an increase in temperature has occurred. Both conduction (via contact) and
convection occur to transfer energy in this system.

46

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

Check This 7.27. Determination of qP(reaction) for dissolution of urea in a calorimeter


The information from Consider This 7.23 indicates that T for the dissolution of urea in
Investigate This 7.22 is 3.5 to 3.6 C. Let us take T = 3.55 C, but remember that we only
really measure the change to two significant figures, so the best we can do is a precision of
about 3% in our result. The simple experimental setup probably makes it even more uncertain,
so we report two significant figures in the result.
qP(solution) = (m)(c)(T) = (106. g)(4.18 Jg1C1)(3.55 C) = 1.6 103 J
Rearrange equation (7.3) to get qP(reaction):
qP(reaction) = qP(solution) = 1.6 103 J
Check This 7.29. Determination of Hreaction for dissolution of urea in a calorimeter
The molar mass of urea, NH2C(O)NH2, is 60 gmol1 and 6.0 g were dissolved in 100. mL of
water in Investigate This 7.22. Thus, 0.10 mol of urea was dissolved and gave a 1.0 M solution.
The molar enthalpy change for the dissolution is:
1.6 10 3 J
Hreaction = qP(reaction)(per mole) =
= 16 kJmol1
0.1 mol
Check This 7.31. Effect of increasing the amount of another calorimetric reaction
(a) In the previous example, Consider This 7.30, we found that doubling the amount of reactants
did not change the temperature change for their reaction, so perhaps it is surprising to find that,
in this case, doubling the amount of urea dissolved in the same volume of water doubles the
temperature change.
(b) The difference between the two examples is that here the system (urea dissolving) was
doubled, but the surroundings (the solution) remained essentially the same. The surroundings
are a little different when double the amount of urea dissolves, because the mass of the solution
and perhaps its specific heat will be a bit different, but these are relatively small changes and we
expect the change in temperature to about double, as it does. In the example in Consider This
7.30, both the reacting system and the surroundings (total amount of solution) were doubled, so
the doubled energy released had to warm twice as much solution.
Check This 7.34 Properties of the possible H2O2 reaction products
(a) Reaction (7.6) produces a gas that consists of oxygen. The gas will support combustion. If a
glowing splint is placed in the gas, the splint will glow more brightly and may burst into flame.
(b) Reaction (7.7) produces a mixture of hydrogen and oxygen gases. This is an explosive
mixture. Hydrogen is flammable and when mixed with oxygen will react so rapidly that an
explosion occurs. If the amount of gas is small the explosion will not be very dangerous and is
likely to produce only a pop sound instead of a deafening blast.
(c) Insert a glowing splint in the cylinder containing the product gas(es).
If reaction (7.6) produces the gas, the glowing splint will glow brightly or even re-ignite.
If reaction (7.7) produces the gas, the glowing splint will ignite the mixture of hydrogen and
oxygen and a small explosion will be heard.
Check This 7.37. Overall homolytic bond cleavage reaction of methane
Each reaction in Table 7.2 produces a carbon-containing species that is the reactant in the next
reaction. If we add the reactions together, these products and reactants cancel, as shown here:

ACS Chemistry FROG

47

Chemical Energetics: Enthalpy

Chapter 7

There are eight bonding electrons in methane, represented as four pairs of bonding electrons by
the strokes in the CH4 structure. The products also contain eight valence electrons, four on the
carbon atom and one each on the four hydrogen atoms, so all the bonding electrons in methane
are accounted for in the products.
Check This 7.39. Formation of polyethylene
The polyethylene "product" should be similar to the structure shown in Figure 7.10, except that
it will contain more monomer, -(-CHCH-)- units.
Check This 7.40. Sulfursulfur bonds
(a) Lewis structures for S8 and S2 are:

(b) A molecular model of the S8 ring (using oxygen atoms) is shown here:

Four S2 models will have two connectors (probably bent) between each pair of sulfurs,
representing a total of eight two-electron (electron pair) bonds. The model here shows that there
are eight two-electron bonds in S8. The eight connectors used for the S2s could have been used
to build the S8. The reactants and products in the model reaction have the same number of twoelectron bonds. We can compare the reactants and products, using bond enthalpies (as in Figure
7.8), because we are comparing species with the same number of bonding electrons.

48

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

(c) A bond formation enthalpy diagram for the reactants, 4S2, and product, S8, of the reaction
represented by the models in part (b) is:

The H for 8S(g) 4S2(g) is 1408 kJmol-1 [= 4(352 kJmol-1)]. The H for 8S(g) S8(g)
is 2144 kJmol-1 [= 8(268 kJmol-1)]. The reaction modeled in part (b) is exothermic: H =
736 kJmol-1.
(d) The results in part (c), based on bond enthalpies, show that the singly-bonded S8 molecule is
a great deal more stable than four doubly-bonded S2 molecules. This is because the double bond
is weaker than two single bonds. The opposite is true for oxygen where a double bond is
stronger than two single bonds in oxygen, so the doubly-bonded species is the more stable.
Thus, a good deal of the difference between the two elements can be attributed to the relative
strengths of their single and multiple bonds.
Check This 7.42. Reaction enthalpy from average bond enthalpies
(a) We can use equation (7.14) to get Hreaction from bond enthalpies (all bond enthalpies in
kJmol1):
Hreaction = (BHreactants) + (-BHproducts)
= BH(OO) + 2BH(OH) BH(O=O) BH(HH)
= 142 + 2(460) 498.7 436.4
= (1062 kJmol1) + (-935 kJmol1) = 127 kJmol1
(b) The enthalpy diagram for this reaction is:

ACS Chemistry FROG

49

Chemical Energetics: Enthalpy

Chapter 7

(c) The interactive exercise in the Web Companion, Chapter 7, Section 7.7, page 3, shows the
cumulative result of breaking all the bonds in the reactants and then making all the bonds in the
products, in order to find the overall Hreaction for the reaction of water with ethene to form
ethanol. This is exactly the procedure we have used in part (a) here and in Worked Example
7.41. Figure 7.11 and our figure in part (b) are just like the enthalpy level diagram in the
Companion, except that our figures do not show the components of diagram, as the Companion
does. On page 2 of this section of the Companion, a shorter method for doing these kinds of
problems is hinted at, but not fully carried out. In this approach, you focus only on the bonds in
the reactants that are broken going from reactants to products and only those bonds in the
product that must be made to create the product from the pieces on the reactant side. The bonds
that remain as they are when going from reactants to products are ignored in bond enthalpy
calculations. For the water-ethene reaction, only the carbon-carbon double bond and one of the
hydrogen-oxygen bonds needs to be broken in the reactants; if you try to break other bonds, you
get this error message:

To get the product, you need to make carbon-carbon, carbon-hydrogen, and carbon-oxygen
single bonds. The bonds that have to be broken in the reactants require a total of 1080 kJmol1.
Making the bonds in the product releases a total of 1112 kJmol1. The reaction is exothermic
and Hreaction = 32 kJmol1. This is the same numerical value you get from the diagram on
page 3 of this section of the Companion, demonstrating that this shorter method gives the same
result as you get by going through the complete atomization. This screen from page 3 illustrates
why the two methods give the same result:

Here you can see that all of the bonds in the reactants have been broken; the first two broken are
the two that must be broken to produce fragments that can be recombined to yield the product.

50

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

In the product, only those bonds that are common to the reactants and products have been
reformed. Note that the enthalpy regained (released) in this reformation is exactly the same as
the enthalpy required for the bond breaking, as it must be. The only bonds remaining to be
formed are those that combine the fragments (H, CH2, CH2, and OH) to give the ethanol
product. These bond enthalpies are the ones to compare to the first two broken to find the net
change for the process. Try this yourself with the Companion to see what you get when the final
three bonds are formed.
Check This 7.43. The hydrogen peroxide decomposition reaction
(a) The enthalpy change as calculated from Check This 7.42 is endothermic. However, the
enthalpy change for this reaction is exothermic, as observed in Investigate This 7.32. The
graduated cylinder felt warm after the reaction.
(b) Reaction (7.6) is responsible for the observations in Investigate This 7.32 and 7.35. The
reaction was exothermic. The glowing splint confirmed the presence of oxygen gas. No small
explosion was detected, so the presence of a mixture of hydrogen and oxygen gases is unlikely.
(The absence of a positive test cannot always be taken as proving the absence of the system
being tested for, because other circumstances might prevent the test working correctly. In this
case, this is one more piece of evidence that is consistent with all the others and strengthens the
final conclusion.) The calculated enthalpy change for reaction (7.6) is exothermic.
Check This 7.45. Formation of compounds from their standard-state elements
The reaction equation for forming dimethyl amine from its elements is:
2C(graphite) + 7/2H2(g) + 1/2N2(g) (CH3)2NH(g)
The reaction equation for forming ethylamine from its elements is:
2C(graphite) + 7/2H2(g) + 1/2N2(g) CH3CH2NH2(g)
The reactants in both examples are the same. We conclude that the reactants in equations for
forming isomers from the elements in their standard states are the same.
Check This 7.47. Standard enthalpy change for an isomerization reaction
(a) The molecular model for the reaction of cyclopentane to form one of its isomers, 1-pentene,
is:

(b) The enthalpy change for this reaction is the difference between the initial and final states,
that is, the enthalpy of formation of the product minus the enthalpy of formation of the reactant:
Hreaction = Hf(1-pentene) Hf(cyclopentane)
= (20.92 kJ mol-1) (77.24 kJ mol-1) = 56.32 kJ mol-1
(c) The enthalpy level diagram (modeled after Figure 7.13) for this reaction is:

ACS Chemistry FROG

51

Chemical Energetics: Enthalpy

Chapter 7

This diagram shows graphically what we found in part (b), the reaction is endothermic.
Check This 7.48. Comparison of formation of 1-butene from its atoms and its elements
(a)

(b) Since cyclo-C4H8 and 1-butene have the same chemical formula, their Hf from graphite
and hydrogen gas will be the same. However, since their molecular structures are different, their
Hreaction will be different.
Check This 7.50
(a) Hreaction = -2801.64 kJmol-1
(b)

52

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

Check This 7.51


(a) Hreaction using bond enthalpies: - 86 kJ mol-1
(b) Hreaction using standard enthalpy for formation data: - 126.3 kJ mol-1
(c) The results are not the same. Bond enthalpies are simply estimates, but are useful in

providing the direction of enthalpy change. Bond enthalpy calculations assume that
when two atoms make a covalent bond, this bond is independent and no other bonds
around it influence it. The data used in bond enthalpy calculations assume that the
products and reactants are in the gas phase. Average bond data are given as a reasonable
approximation, again independent of the molecule. For example, although ethene and
1,2-dibromoethane are two different molecules, the same C-H bond enthalpy data was
used in the calculations. The stand enthalpies of formation are more accurate because
chemical bonding considerations are not involved in these calculations.
Check This 7.52
The bond that links the first phosphate group to deoxyribose is a phosphate ester.

ACS Chemistry FROG

53

Chemical Energetics: Enthalpy

Chapter 7

Check This 7.53


Although the fats in the corn chips and the sugars in the marshmallow seem to provide an initial

burst of energy, the spaghetti and pecan provide energy at a slower, steady rate. Thus, this type
of fuel seems to be the best because it can provide energy over a length of time. Corn chips and
marshmallows seems to provide an initial burst of energy that will drop off quickly.
Check This 7.54
(a) The advance to the organism is twice the free energy of hydrolysis of ATP alone, so the dual

hydrolysis can be coupled with reactions that require more driving force than furnished by the
single hydrolysis. (b) The phosphoric anhydride bonds are hydrolyzed in each case, so the
enthalpy released is approximately the same.
Check This 7.57
w = (-103 kPa)(-0.521 L) = 53.7 kJ
Check This 7.60
E and H are the same in reactions (b) and (d). In reaction (b), V = 0. Reaction (d) involves

no gases in reactants or products. Therefore, V = 0 as well. In reaction (a), V is negative and


-PV is positive. Work is done on the system in this reaction. In reaction (c), V is positive and
-PV is negative. Work is done on the surroundings.
Check This 7.62
(a) E = qp + wp = Hreaction + wp = 126.3 kJ + 2.5 kJ = 128.8 kJ
(b) For this case, we see that Horeaction and Eoreaction differ by a little less than 2%. Thus, the

approximation that Eoreaction Horeaction is quite good, but not as good as when the thermal
change in the reaction was much larger. You can see from Worked Example 7.61 and this Check
This that, for a one mole change in the amount of gaseous products vs. reactants, the work done
by or on the system under standard conditions at 1 bar (= 100 kPa) pressure is 2.5 kJ. This is a
rather small amount of energy that will be swamped by large thermal changes for most reactions.
The approximation that Eoreaction Horeaction breaks down when the thermal change in the
reaction is relatively small, say no more than 50 kJ.
Check This 7.67
Pressure is inversely proportional to volume. Pressure is directly proportional to temperature and

the number of molecules.


54

ACS Chemistry FROG

Chapter 7

Chemical Energetics: Enthalpy

Check This 7.69


PiVi = PfVf; Vf =

(100. kPa )(44.7 mL)


150 kPa

= 29.8 mL

Check This 7.71


(a) In this problem, the volume remains the same while n2 moles of a second gas are added to n1
moles of the first gas. In the process, the pressure and temperature of the gas sample changes.
We know the initial conditions, P1, T1, n = n1, and the final conditions, P2, T2, n = (n1 + n2), so
we can write:
R
P
P2
= 1 =
V n1T1 (n1 + n2 )T2

57.8 kPa
95.8 kPa
=
n1 289.2 K (n1 + n2 ) 302.7 K
(b) n =

PV
(95.8 kPa)(0.547 L))
=
= 0.021 moles of gas
RT (8.314 J mol-1 K-1)(302.7 K)

Check This 7.77


H = qP = 1,129 J / 8.1 X 10-2 mol = 13.9 kJ mol-1

E = 961 J / 8.1 X 10-2 mol = 11.9 kJ mol-1

ACS Chemistry FROG

55

You might also like