Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Chemistry: With Inorganic Qualitative Analysis
Chemistry: With Inorganic Qualitative Analysis
Chemistry: With Inorganic Qualitative Analysis
Ebook2,870 pages

Chemistry: With Inorganic Qualitative Analysis

Rating: 5 out of 5 stars

5/5

()

Read preview

About this ebook

Chemistry with Inorganic Qualitative Analysis is a textbook that describes the application of the principles of equilibrium represented in qualitative analysis and the properties of ions arising from the reactions of the analysis. This book reviews the chemistry of inorganic substances as the science of matter, the units of measure used, atoms, atomic structure, thermochemistry, nuclear chemistry, molecules, and ions in action. This text also describes the chemical bonds, the representative elements, the changes of state, water and the hydrosphere (which also covers water pollution and water purification). Water purification occurs in nature through the usual water cycle and by the action of microorganisms. The air flushes dissolved gases and volatile pollutants; when water seeps through the soil, it filters solids as they settle in the bottom of placid lakes. Microorganisms break down large organic molecules containing mostly carbon, hydrogen, nitrogen, oxygen, sulfur, or phosphorus into harmless molecules and ions. This text notes that natural purification occurs if the level of contaminants is not so excessive. This textbook is suitable for both chemistry teachers and students.
LanguageEnglish
Release dateDec 2, 2012
ISBN9780323141758
Chemistry: With Inorganic Qualitative Analysis

Related to Chemistry

Chemistry For You

View More

Reviews for Chemistry

Rating: 5 out of 5 stars
5/5

2 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Chemistry - Therald Moeller

    METZ

    1

    CHEMISTRY: THE SCIENCE OF MATTER

    Publisher Summary

    This chapter presents some basic definitions of science, matter, and chemistry. It explains the subdivisions of chemistry. The SI system defines a single base unit to measure each of seven quantities such as length, mass, time, electric current, temperature, amount of substance, and luminous intensity. Other necessary units, such as that for volume, are derived from the base units. The chapter reviews the modern units of measurement and a problem-solving method, the factor-dimensional method. With the factor-dimensional method, a unit-conversion calculation, and the calculation needed to solve a problem, can be set up in a single expression. The chapter describes the role of chemistry in coping with the major problems facing mankind.

    In this chapter we first give some basic definitions—science, matter, chemistry—and then an explanation of the subdivisions of chemistry. We review the modern units of measurement and a problem-solving method—the factor-dimensional method—that will be used throughout this book. In the last section of the chapter we take a brief look at the role of chemistry in coping with the major problems facing mankind.

    Look around you. That’s how chemistry began—in the limitless curiosity of human beings about their surroundings.

    Possibly you are sitting at your desk with some paper and a wooden pencil or a plastic pen at hand to take notes. Maybe there are some metal paper clips and a pottery coffee cup or a glass soft-drink bottle or an aluminum can on your desk. What could you do to investigate the materials in your paper, your pencil or pen, the paper clips, the cup, the bottle, or the can? Scratch them. Which is harder? Put a drop of water, or alcohol, or acid on each one. What happens? Weigh pieces of equal size. Which is heavier? Try to burn a small piece of each. Which ones burn? What is left afterwards?

    You could work your way around your room cataloging how everything in it responds to these and other tests. You could go outside and do the same for the rocks and plants. Pretty soon you would be able to draw conclusions about which things are similar to each other and which differ from each other.

    If you have a curious nature, your next questions should begin with Why…? and How…? Why does wood burn, but pottery not? How can I predict whether other things will or won’t bum? Why are some things heavier or harder than others? Why don’t these things dissolve in water? How does acid change a paper clip?

    Chemistry has its roots in just this kind of speculation about the nature of simple things. In early times, people wondered about air, and water, and rocks, and fire, and looked for magical and mystical answers to questions about the physical world around them. Once the importance of systematic observation was recognized, the foundation was laid for chemistry and all the sciences.

    Science and matter

    1.1 Science

    A natural science is a classification of knowledge about things that are observable in nature, in the material world, and in the universe. Each branch of science organizes a multitude of facts and answers to How…? and Why…? questions. In the biological sciences, questions are asked mainly about things that are alive. In the physical sciences, the questions pertain mainly to things that are not alive. Botany and zoology are biological sciences, geology and meteorology are physical sciences.

    Chemistry is in a central position. It applies mathematics and the laws of physics, and is thought of most often as a physical science. But the chemical elements are the building blocks for everything in the universe, living or not. Chemistry, therefore, legitimately asks questions about life itself.

    The boundaries between the biological sciences and chemistry are becoming increasingly blurred. These boundaries will continue to fade as scientists solve such puzzles as the structure of the genes that carry the message of life to future generations. But despite such changes, the scientific method will remain, for all of science is built upon its principles. The scientific approach to solving a problem frequently begins with systematic observations. Once enough observations have been made a hypothesis can be formulated. At this stage a hypothesis is a tentative explanation for a set of observations.

    Experiments are next devised to test the validity of the hypothesis in as many ways as possible. If a hypothesis survives the test of many experiments, it may be accepted as a law. A law is a statement of a relation between phenomena that, so far as is known, is always the same under the same conditions. As knowledge in a particular area of science grows, further experiments may lead to further laws. Eventually it may be possible to formulate a theory. Theories have larger scope than laws. A theory is a unifying principle or group of principles that explains a body of facts or phenomena and those laws that are based upon them.

    Central to the work of scientists and to progress in science is a constant testing of current assumptions as well as a constant search for answers to new problems. However, to talk of the scientific method may lead you to believe that scientific inquiry always moves forward in a rigid, step-by-step progression. Seldom is this true. Sometimes law precedes theory, sometimes theory precedes law. Facts and observations may come first, or a purely theoretical idea may come first. Usually everything is happening at once.

    1.2 States and properties of matter

    Almost all definitions of chemistry contain the word matter or the word materials. Matter is all around us. Matter is everything of which the world is made. It occupies space, it has mass, and, except for some gases, it can be seen and touched.

    There are three states of matter—the gaseous state, the liquid state, and the solid state. Some of the general properties that distinguish gases, liquids, and solids from each other are listed in Table 1.1.

    TABLE 1.1

    General properties of the gaseous, liquid, and solid states

    aCompressibility is a decrease in volume with increase of pressure.

    bExpandability is an increase in volume with increase of temperature.

    cViscosity is the resistance of a substance to flow.

    At ordinary temperatures and pressures oxygen, nitrogen, hydrogen, carbon dioxide, chlorine, ammonia, and methane (the major component of natural gas) are all gases. Water, ethyl alcohol, mercury, and gasoline are liquids at ordinary temperatures and pressures. And solids, of course, are everywhere we look, most of them combinations of several kinds of substances. Some common simple substances are solids at ordinary temperatures and pressures, including most metals, such as iron, copper, and gold; carbon, either as diamond or graphite; sodium chloride (common salt); and sucrose (common sugar). But pottery, the substance from which your coffee cup is made, is a mixture of complex materials that contain silicon, aluminum, oxygen, and small amounts of various metals.

    Some substances can exist in all three states. Water is known in the solid state as ice, in the liquid state (at room temperature), and in the gaseous state as steam. Many metals, which are usually solid, can be melted, and if heated to higher temperatures, can become gaseous. Some substances, however, cannot exist in the gaseous state; others cannot exist in the liquid state, and some cannot exist in either the gaseous or the liquid state. For example, calcium carbonate, a solid, cannot be melted or vaporized, for upon heating it decomposes into calcium oxide, which is another solid, and carbon dioxide. Upon gentle heating, sugar melts to the liquid state. Upon heating to higher temperatures sugar does not become gaseous, but instead decomposes into a variety of products that contain carbon. However, all gases and liquids can be condensed to the solid state.

    The careful examination and measurement of the properties of different substances are important in chemistry. Color, melting point, boiling point, density, and hardness are among the observable properties of matter. These are physical properties. Physical properties can be exhibited, measured, or observed without resulting in a change in the composition and identity of a substance. Chemical properties, by contrast, can only be observed in chemical reactions. A chemical reaction is a process in which at least one substance is changed in composition and identity. The burning of wood in air and the tarnishing of silver demonstrate chemical properties. The burning wood is combining with oxygen in the air to form carbon dioxide, water, and carbon-containing substances, which we see as smoke and ashes. The shiny silver is combining with sulfides in the air to form silver sulfide, the black substance that stains the surface. Before we can define chemistry further, however, we must distinguish between the different kinds of matter that we will encounter.

    1.3 Kinds of matter

    Every kind of matter can be classified as either a pure substance or a mixture. A pure substance is a form of matter that has identical physical and chemical properties no matter what its source. For instance, pure water is colorless and odorless, boils at 100°C and freezes at 0°C at atmospheric pressure, weighs 1 gram per milliliter at 4°C, and does not burn. It has these properties whether it is distilled from sea water, dipped from a mountain stream, or prepared in a chemical reaction by the union of hydrogen and oxygen. And it is these characteristic properties that enable us to distinguish water from other substances.

    A mixture is any combination of two or more substances in which the substances combined retain their identity. The substances in a mixture can be combined in any proportions. The properties of the mixture are the properties of its components, and the components can be retrieved intact from the mixture without a chemical reaction. A heterogeneous mixture is a mixture in which the individual components of the mixture remain physically separate and can be seen as separate components, although in some cases a microscope is needed. Concrete and granite are heterogeneous mixtures; so are milk and cake batter. Powdered iron and powdered sulfur, no matter how well stirred, form a heterogeneous mixture. They can be separated by using a magnet to attract the powdered iron, as shown in Figure 1.1a.

    FIGURE 1.1 Iron plus sulfur: a mixture or a chemical compound. If iron and sulfur powders are thoroughly mixed but not heated, a mixture results. The iron can be completely removed from the sulfur by a magnet. Conversely, the sulfur can be completely removed from the iron by dissolving it in carbon disulfude, a liquid in which iron is not soluble. Powdered iron and powdered sulfur when heated together give a chemical compound, iron sulfide.

    The substances in a homogeneous mixture are thoroughly intermingled; the composition and appearance of the mixture are uniform throughout. Air is a homogeneous mixture of gases, and motor oil is a homogeneous mixture of liquid petroleum derivatives. Homogeneous mixtures, whether gaseous, liquid, or solid, also can be separated by physical means.

    Pure substances can be either elements or compounds. An element is a pure substance that cannot be converted into a simpler form of matter by any chemical reaction. Hydrogen and oxygen are elements; so are iron and sulfur. All of the known elements are listed on the endpapers of this book.

    Under the right conditions hydrogen and oxygen combine to form water. Heating a mixture of iron and sulfur causes them to combine to form iron sulfide, a substance with properties clearly different from those of either iron or sulfur (Figure 1.1b). Both of these processes are chemical reactions, and the products are chemical compounds. A compound is a substance of definite composition in which two or more elements are chemically combined; a compound can be separated into its components by chemical reactions but not by purely physical methods.

    The general classification of matter is summarized in Figure 1.2. We shall define the types of matter more accurately in later chapters.

    FIGURE 1.2 Classification of matter. By naturally occurring elements, we mean elements that are present in nature in even the minutest amounts. Four elements—technetium, astatine, francium, and promethium—although present, are so scarce that collecting a measurable amount is exceedingly difficult or impossible.

    Chemistry: The science of matter

    1.4 Chemistry

    Simply defined, chemistry is the science of matter. Chemistry is necessary in the study and manipulation of any material, and in our modern world we are surrounded by materials that have been studied, manipulated, and even invented by chemistry. Our automobiles are made of metals, fabrics, and plastics. Few of these materials can be obtained from natural sources in forms that are readily usable. The metals are recovered from mineral ores; the fabrics are produced from plant, animal, or synthetic fibers; the plastics are made by combining simple materials to form new and more complex materials.

    We travel on highways surfaced with asphalt and concrete. We drink water that has been purified with chlorine. We read from paper made from cellulose and printed with highly colored inks. We watch television screens that have been coated with tiny chemical spots, each of which responds with a color when energy is applied. With chemistry, we alter an almost endless variety of materials to form other materials with characteristics that we desire.

    Not all chemistry has a practical product as its objective, of course. Chemists search for understanding and the ability, based on understanding, to predict what will happen when changes occur. And some chemists pursue inquiries into the nature of matter solely because they enjoy it.

    With these observations in mind, we may consider a more formal definition of chemistry: Chemistry is the branch of science that deals with matter, with the changes that matter can undergo, and with the laws that describe these changes. As this definition implies, chemistry is both a theoretical and an applied science. The principles of chemistry are the explanations of the chemical facts; this is where you meet the hypotheses, the laws, and the theories. Descriptive chemistry, as you might expect, is the description of the elements and compounds, their physical states, and how they behave. No matter how chemistry is used, a good balance between principles and descriptive chemistry is a necessity. We will attempt to maintain such a balance throughout this textbook.

    1.5 Subdivisions of chemistry

    The five major subdivisions of chemistry are listed in Table 1.2. Originally, organic chemistry dealt only with substances obtained from living materials, but this distinction has long since vanished. Organic chemistry has become the chemistry of carbon compounds and their derivatives. Organic compounds that contain only the two elements carbon and hydrogen are called hydrocarbons. Almost all other organic compounds can be thought of as derivatives of hydrocarbons. Inorganic chemistry is the chemistry of all elements (including carbon) and their compounds, with the exception of hydrocarbons and hydrocarbon derivatives. The other major subdivisions—analytical chemistry, physical chemistry, biochemistry—are defined briefly in Table 1.2.

    TABLE 1.2

    Major subdivisions of chemistry

    The boundaries between organic and inorganic chemistry, like the boundaries between the different sciences, and indeed between all areas of chemistry, are disappearing. Many compounds between organic and metallic substances—organometallic compounds—are now known, and organic compounds have been made that behave in ways traditionally attributed only to inorganic compounds.

    Some specialized areas of chemistry, which provide bridges to other major fields of study and endeavor, are geochemistry; microbial and medicinal chemistry; agricultural, fertilizer, soil, and food chemistry; polymer chemistry; cellulose, paper, and textile chemistry; and industrial and environmental chemistry. As we have pointed out, whenever people explore materials, living or nonliving, they eventually become concerned with chemistry.

    Units of measure; problem solving

    1.6 Systems of measurement

    Since 1793, when the metric system was devised by the French National Academy of Sciences to overcome the profusion of units handed down from medieval times, various international bodies have been defining and redefining units of measurement and attempting to gain widespread uniformity in their use. For many years scientists everywhere and people in most European countries have used the metric system, which is a decimal system.

    In 1960, the International Bureau of Weights and Measures adopted the International System of Units, known as SI units (for Système Internationale). The SI system is a revision and extension of the metric system; it provides basic units for each type of measurement. Scientists and engineers throughout the world in all disciplines are now being urged to use the SI system exclusively; most major countries have adopted it.

    The United States, until 1975, officially stayed with a weights and measures system based upon the English system of inches and feet, ounces and pounds, pints and quarts, and so on. The Metric Conversion Act of 1975 commits this country to a policy of voluntary conversion to the metric system by 1985. The United States Metric Board has the job of coordinating the changeover and educating the public to the use of the new units. Everyone must learn how to buy a liter of milk and what to wear when the temperature is 4°C.

    The kilogram and the liter are standard units in both the original metric system and the new SI system, and scientists have dealt in kilograms and liters for a long time. However, some of the units in the SI system are supposed to replace other units that are still widely used. The SI system is finding acceptance, but because scientists are human and, like everyone else, resistant to changing their ways, it will undoubtedly be a long time before the changeover to the new system is complete.

    The SI system defines a single base unit to measure each of seven quantities—length, mass, time, electric current, temperature, amount of substance, and luminous intensity (see Table 1.3). Other necessary units, such as that for volume, are derived from the base units. A standard set of prefixes indicates variations in the magnitude of these units, as shown in Table 1.4. Some equivalents between SI units and other units are given in Table 1.5. (See also Appendix B.)

    TABLE 1.3

    SI base units

    TABLE 1.4

    Unit prefixes for SI units

    The prefix is written in front of the unit name with no hyphen, e.g., micrometer, nanometer, or the symbol is written in front of the unit abbreviation, e.g., μm (micrometer), nm (nanometer); μ is the lower case Greek letter mu.

    TABLE 1.5

    Equivalence between units

    Length

    1 inch = 2.54 cm

    1 ft = 30.48 cm

    1 yd = 0.9144 m

    1 mi = 1.609 km

    1 cm = 0.3937 inch

    1 m = 39.37 inches

    1 Å = 0.1 nm

    Volume (liquid)

    1 qt = 0.946 liter

    1 gal = 3.785 liters

    Weight, mass

    1 oz = 28.35 g

    1 lb = 453.6 g

    1 g = 2.205 × 10−3 lb

    1 kg = 2.205 lb

    Heat, energy

    1 cal = 4.184 J

    1 erg = 1 × 10−7 J

    1 erg = 2.39 × 10−8 cal

    The fractions and multiples represented by the prefixes in Table 1.4 and some of the unit equivalents in Table 1.5 have been written as exponential numbers. Very large and very small numbers arise often in chemistry. Such numbers are easier to handle and easier to comprehend when they are given in exponential form. If you do not know how to express numbers in exponential form, also called scientific notation, you should study Appendix A.2 at this point.

    In the following sections, we discuss the units encountered most often in chemistry.

    1.7 Length

    In both the metric and SI systems, the meter is the basic unit of length or distance. Chemists often have to talk about very small lengths, such as the dimensions of particles of matter. In this book, we use the nanometer (nm), an SI unit, for this purpose. A nanometer is one billionth of a meter, or 1 × 10−9 m.

    1.8 Volume

    The milliliter and the liter are the standard units of volume in most chemical laboratory work. One milliliter is exactly equal to one cubic centimeter (cm³), and one liter exactly equals 1000 milliliters (ml) or 1000 cubic centimeters. The SI system recommends replacing the liter with the cubic decimeter, but in this book we have chosen to use milliliters and liters. (From 1901 to 1964, a liter was defined as the volume of 1 kilogram of water at 4°C. During this period a milliliter was very slightly larger than a cubic centimeter. In 1964, the liter was redefined as exactly equal in volume to 1000 cubic centimeters, thereby getting rid of some confusion.)

    1.9 Mass vs. weight

    The distinction between mass and weight should be clear to anyone who watched the astronauts bounding over the surface of the moon. The moon’s gravity is smaller than that of the Earth, and so the weight of the astronauts was less there. Their bodies, however, were unchanged and had the same mass as always. Mass is an intrinsic property and represents the quantity of matter in a body. Weight is the force a body exerts because of the pull of gravity on the body’s mass. Now that we have carefully made this distinction, we must admit that in talking and writing about chemistry, the distinction is often ignored, and the terms mass and weight are used imprecisely.

    Both the SI and metric systems rely on the gram and its fractions and multiples for units of mass, with the kilogram designated the basic mass unit by the SI system.

    Mass and volume are properties that by themselves disclose nothing about the identity of a substance. But when they are combined as mass per unit volume, called density, these units give one of those properties which does describe a substance. For example, the density of aluminum is 2.7 g/cm³, the density of iron is 7.8 g/cm³, and the density of mercury is 13.6 g/cm³; osmium and iridium, the densest of the elements, have densities of 22.61 and 22.65 g/cm³, respectively.

    1.10 Heat

    When a warm body and a cold body come into contact, heat energy (thermal energy) flows from the body with the higher temperature to that with the lower temperature. Heat is given off or taken up in almost all chemical changes and in many physical changes. The calorie (cal) is a unit of measure of heat. One calorie is close to the amount of heat needed to raise the temperature of 1 gram of water by 1°C. Because the magnitude of the numbers is more convenient, the heat in chemical changes is usually expressed in kilocalories (kcal). The calories counted by dieters are really kilocalories.

    In the SI system, the basic unit of energy is the joule (J). In this book, you will be given the opportunity to use both kilojoules and kilocalories, because kilocalories are still in widespread use. The units of length, mass, volume, and heat used in chemistry are summarized in Table 1.6.

    TABLE 1.6

    Units in chemistry

    Included here are the units most often encountered in chemistry. Those given in color are used throughout this book. Those given in black may be encountered in other chemistry books or in your study of other sciences.

    1.11 Temperature

    You should be familiar with three temperature scales: the Kelvin scale, °K; the Celsius scale, °C; and the Fahrenheit scale, °F (see Figure 1.3). On the Fahrenheit scale, the one that has been in everyday use in the United States, the freezing point of water is 32°, the boiling point of water (at the average atmospheric pressure at sea level) is 212°, and the temperature range between these two points is divided into 180 equal degrees. The Celsius and Fahrenheit scales coincide at -40° (see Figure 1.3).

    FIGURE 1.3 Comparison of °C,°F, and °K.

    The Celsius scale has been used in most scientific work; its reference points are 0° for the freezing point of water and 100° for the boiling point of water (at the average atmospheric pressure at sea level). Since the same temperature range is represented by 180° on the Fahrenheit scale and 100° on the Celsius scale, a Fahrenheit degree is 100/180 or 5/9 of a Celsius degree. Anders Celsius described a thermometer using this scale in 1742. The same scale has been called the centigrade scale. Fortunately, both are represented by the symbol, °C.

    Temperatures on the Celsius and Fahrenheit scales can be inter-converted by the following equations (Note the difference in placement of the parentheses.)

    The Kelvin temperature unit is one of the basic SI units. On the Kelvin scale, 0° represents absolute zero, the lowest temperature attainable. The freezing point of water is 273.15°K and the boiling point is 373.15°K. Therefore, the Kelvin degree and the Celsius degree are the same size, and a Celsius temperature can be converted to a Kelvin temperature by adding 273.15. For general purposes, 273° is sufficiently accurate:

    Although SI recommends the Kelvin, K, without a degree sign, we have chosen to use °K.

    1.12 The factor-dimensional method of calculation

    The factor-dimensional method of calculation is a powerful and useful technique for solving problems. Units are retained with all numbers in setting up the solution to a problem; then the units are canceled in the same manner as numbers are canceled. With this method, a seemingly complicated problem can be reasoned out step by step. This process leads to a much better understanding than if the problem were solved with some memorized formulas. If a problem is set up correctly, the answer will be in the correct units. A wrong method of solving the problem is instantly recognizable, for the answer will be in the wrong units.

    In this section, some unit-conversion problems are solved by the factor-dimensional method. The principles remain the same when the method is used for solving many other kinds of problems in chemistry.

    If a measurement in centimeters—say, 94.0 cm—must be converted to meters, the factor-dimensional equation looks like this:

    The conversion factor is obtained from the relationship 1 m = 100 cm by noting that this expression means that there is 1 m per 100 cm, or 1 m/100 cm. To convert from meters to centimeters, the conversion factor would be 100 cm/1 m.

    Here is a more complicated example:

    EXAMPLE 1.1

    Convert 16.1 km to miles using the following unit equivalents:

    1 km = 1000 m

    1 m = 100 cm

    1 inch = 2.54 cm

    1 ft = 12 inches

    1 mi = 5280 ft

    The calculation is set up as follows:

    In each multiplication, a unit from the preceding step is canceled and replaced by a new unit. Note that the answer to this problem, 10.0 mi, is given to three places, that is, as 10.0 rather than 10 or 10.00. The choice of how many digits to give in an answer is governed by the rules of significant figures, which reflect the precision of measurements. In these days when nearly everyone has a pocket calculator, knowing how many of the eight or ten digits that appear in an answer are significant is very important. If you are not familiar with the concept of significant figures, you should study Appendix A.1.

    The factor-dimensional method of calculation can be summarized in four steps:

    1. Determine what is to be calculated and what its units should be.

    2. Using the data in the problem and any necessary conversion factors, set up the calculation.

    3. Include units in each factor and cancel units just as you would numbers.

    4. Write the numerical answer with its units. If the units obtained are incorrect, recheck the conversion factors and the cancellations.

    As Example 1.2 shows, it is sometimes necessary to square or raise to another power one or more of the conversion factors.

    EXAMPLE 1.2

    Engine capacities in American cars are usually given in cubic-inch displacement; for European cars they are usually given in cubic-centimeter or liter displacement. (Displacement is the total volume swept by the pistons in one stroke.)

    A typical European luxury sports car might have an engine displacement of 3.20 liters. How does this compare with the displacement of a typical American family car, which is about 400 cubic inches?

    Unit equivalents:

    1 inch = 2.54 cm 1.000 cm³ = 1.000 ml 1000 ml = 1.000 liter

    1 cu inch = (2.54)³ cm³

    In order to make the comparison, convert 400 cubic inches to liters:

    A typical American family car has a displacement about twice that of a European luxury sports car.

    With the factor-dimensional method, a unit-conversion calculation and the calculation needed to solve a problem can be set up in a single expression, as in Example 1.3.

    EXAMPLE 1.3

    What is the weight in pounds of a gold bar 12.0 inches long, 6.00 inches wide, and 3.00 inches thick? The density of gold is 19.3 g/cm³.

    Unit equivalents:

    Note that the first two steps in this calculation convert density in g/cm³ to density in lb/cu inch.

    Chemistry and the future

    The number of people living on Earth has been growing at the rate of 2% per year. If this rate of growth continues, the world population will double every 35 years. By 2000 a.d., between six and eight billion people will be living here (Table 1.7). All of them will need food, clothing, and shelter. How much each man, woman, and child can have is going to depend on how well we learn to manage the available natural resources. Highly industralized countries like the United States, England, and Japan place great demands on natural resources in order to provide materials to keep the machines of industry at work. In the United States in 1970, 6% of the world’s population used between one-third and one-half of all the energy supplies and natural resources consumed that year.

    TABLE 1.7

    World population

    Most products of our industrial civilization wind up as trash sooner or later. Gaseous wastes drift out of our smokestacks and exhaust pipes and linger in the air over our cities. Liquid and solid trash is dumped into our streams, rivers, and oceans, where it is distributed around the world. We have been very successful in producing new materials for our industrial products. As a result, much of our trash will be with us far into the future. Aluminum cans, pesticides, and some plastics, for example, do not decay as readily as do the natural materials they have replaced.

    The demands for clean air and water, the resulting environmental protection laws, and the economics of raw-material availability are forcing manufacturers to look anew at how their products are produced and used, and where they wind up after they are used. The energy problem increasingly affects decisions on how to extract raw materials, how to process them, and what products to manufacture. A balance will often have to be struck between the desire to recycle and, thereby, avoid waste and depletion of resources, and the cost in energy of transporting and processing waste for recycling.

    Chemistry has an important role to play in determining what happens. Energy is produced in chemical reactions. Chemistry can increase the yield of crops, prolong the storage life of food, and make protein from petroleum. Chemical methods of birth control can help limit the number of people waiting to eat the food. Manufacturers can use raw materials more efficiently when they better understand the properties of materials, and consumers also can benefit from such an understanding. Chemistry will be needed to keep our water drinkable and our air breathable, and an understanding of chemistry can help government officials and voters see to it that beneficial laws are passed.

    We are not about to tell you that science, and chemistry in particular, is mankind’s great hope. The hard questions about population, food and energy, natural resources, and pollution of our environment will not be answered by chemists, engineers, politicians, economists, sociologists, or any single group. But as you study chemistry, keep in mind that which is sometimes easy to forget—that chemistry has great relevance to the world outside the laboratory and the classroom.

    THOUGHTS ON CHEMISTRY

    Spaceship Earth

    Now there is one outstandingly important fact regarding Spaceship Earth, and that is that no instruction book came with it. I think it very significant that there is no instruction for successfully operating our ship. In view of the infinite attention to all other details displayed by our ship, it must be taken as deliberate and purposeful that an instruction book was omitted. Lack of instructions has forced us to find that there are two kinds of berries—red berries that kill us and red berries that will nourish us. And we had to find out ways of telling which-was-which red berry before we ate it or otherwise we would die. So we were forced, because of the lack of an instruction book, to use our intellect, which is our supreme faculty, to devise scientific experimental procedures and interpret effectively the significance of the experimental findings. Thus, because the instruction manual was missing we are learning how we safely can anticipate the consequences of an increasing number of alternative ways of extending our satisfactory survival and growth—both physical and metaphysical.

    In organizing our grand strategy we must first discover where we are now; that is, what our present navigational position in

    OPERATING MANUAL FOR SPACESHIP EARTH by R. Buckminster Fuller

    the universal scheme of evolution is. To begin our position-fixing aboard our Spaceship Earth we must first acknowledge that the abundance of immediately consumable, obviously desirable or utterly essential resources have been sufficient until now to allow us to carry on in spite of our ignorance. Being eventually exhaustable and spoilable, they have been adequate only up to this critical moment. This cushion-for-error of humanity’s survival and growth up to now was apparently provided just as a bird inside of the egg is provided with liquid nutriment to develop to a certain point. But then by design the nutriment is exhausted at just the time when the chick is large enough to be able to locomote on its own legs. And so as the chick pecks at the shell seeking more nutriment it inadvertently breaks open the shell. Stepping forth from its initial sanctuary, the young bird must now forage on its own legs and wings to discover the next phase of its regenerative sustenance.

    My own picture of humanity today finds us just about to step out from amongst the pieces of our just one-second-ago broken eggshell. Our innocent, trial-and-error-sustaining nutriment is exhausted. We are faced with an entirely new relationship to the universe. We are going to have to spread our wings of intellect and fly or perish; that is, we must dare immediately to fly by the generalized principles governing the universe and not by the ground rules of yesterday’s superstitious and erroneously conditioned reflexes. And as we attempt competent thinking we immediately begin to reemploy our innate drive for comprehensive understanding.

    operating manual for spaceship earth, by R. Buckminster Fuller, 1970 (pp. 47 and 51)

    Significant terms defined in this chapter

    Note: Significant terms are listed in the order in which they appear in the chapter.

    Exercises

    Note: One asterisk before a problem indicates it is challenging; two asterisks indicate it is a brain teaser.

    1.1. A chemical magician sets four beakers in front of you and announces that all four contain water. A small amount of the liquid from the fourth beaker is poured into the first beaker, producing a red liquid; into the second beaker, producing a white liquid; and into the third beaker, producing a blue liquid. You decide that this demonstration must be a trick and ask to investigate both the original liquids and the final solutions. You find that the original liquid in the first beaker has the obvious proper ties of water, of being colorless and clear. Foolishly, you drink it—a dangerous thing to do in chemistry. You ob serve that it is an excellent laxative. You test a sample of the original liquid in the second beaker and find that the liquid contains water and the compound lead nitrate. Upon careful observation, you notice that the third original liquid has a light blue color characteristic of a solution of the compound cupric nitrate in water. The fourth liquid is slippery to the touch—another foolish test in chemistry—and produces skin burns rather quickly. A deep whiff—another unwise move in chemistry—of the air above the liquid readily identifies it as ammonia in water. The red and blue liquids are quite stable, but the white liquid eventually separates into a white solid on the bottom of the beaker with a clear, colorless liquid above it. Your investigations prove that you were correct in assuming the demonstration to be a trick.

    You have applied the scientific method to this problem. What was your (a) hypothesis, (b) experiment, and (c) theory? List the (d) physical and (e) chemical properties mentioned for the various liquids. Which beakers contained, either before or after mixing, (f) matter, (g) mixtures, (h) homogeneous mixtures, (i) heterogeneous mixtures, (j) only a pure substance, (k) only a pure compound, and (l) only an element?

    1.2. Which of the following types of chemist—(i) analytical chemist, (ii) biochemist, (iii) inorganic chemist, (iv) organic chemist, or (v) physical chemist—would primarily be interested in (a) the preparation of a compound containing carbon, hydrogen, and nitrogen; (b) the determination of the exact formula of the compound; (c) the behavior of the compound in biological systems; (d) the preparation of a new compound containing the carbon-hydrogen-nitrogen compound combined with copper and chlorine; (e) the determination of the exact molecular structure of these compounds? This exercise illustrates the overlap and teamwork between the various disciplines of chemistry.

    1.3. Which statements are true? Rewrite any false statement so that it is correct.

    (a) The individual components in a homogeneous mixture remain physically separate and can be seen as separate components.

    (b) A generalization about various facts or observations is known as a hypothesis.

    (c) The density of a substance depends on the total amount of substance present.

    (d) A theory is a single statement or a series of statements that are so well documented that no experiment is likely to yield data that will refute it.

    (e) Chemistry is the branch of science that deals with matter, with the changes that matter can undergo, and with the laws that describe these changes.

    1.4. Which of the following metric units are used to express (a) mass, (b) energy, (c) length, (d) volume, and (e) temperature: (i) erg, (ii) cm, (iii) cm³, (iv) °K, (v) Å, (vi) J, (vii) km, (viii) ml, (ix) g, (x) °C, (xi) nm, (xii) liter, (xiii) cal, (xiv) m, (xv) dm³, (xvi) kg, and (xvii) mg?

    1.5. Which of the following units are correct for the property that is being measured: (a) the area of a football field in m², (b) the volume of an apple juice bottle in cu liters, (c) the density of wood in kg/m³, (d) the length of an eraser in ml, (e) the radius of a basketball in kg, (f) the length of time of a TV commercial in Msec, and (g) the height of an evergreen tree in cm³?

    1.6. Which of the following conversions are incorrect: (a) 1 m = 2.54 cm, (b) 1000 m = 1 km, (c) 1 ml = 1 cm³, (d) 454 lb = 1 g, (e) 1 Å = 0.1 nm, (f) 1 dm³ = 1 liter, (g) 1 g/cm³ = 1 × 10³ kg/m³, (h) 1°F = 1°K, (i) 1 cal = 4.184 J, (j) 1 J = 1 × 10−7 erg, and (k) 1 mμ = 1 nm?

    1.7. Clearly distinguish between (a) physical and chemical properties, (b) mass and weight, (c) substances and mixtures, (d) elements and compounds, and (e) homogeneous and heterogeneous mixtures.

    1.8. In which of these processes is a chemical reaction taking place: (a) magnetizing a nail, (b) fermenting grape juice, (c) burning of gasoline, (d) cutting meat, (e) cooking meat, (f) baking a cake, (g) turning of leaves in autumn, (h) setting of concrete, and (j) inflating a balloon?

    1.9. Using the factor-dimensional method of calculation, make each of the following conversions: (a) 100.0 yd to cm, (b) 20.0 liter to gal, (c) 1.00 ft³ to mm³, (d) 1.429 g/liter to lb/ft³, (e) 45 mi/hr to m/sec, (f) 440 yd to m, (g) 7 lb 10 oz to kg, (h) 25 gal to liters, (i) 0.0025 inch to nm, and (j) 14300 cal to J.

    1.10. For each case, which is greater: (a) 1.0 cm or 1.0 inch, (b) 35°C or 70.0°F, (c) 300°K or 85°F, (d) 1.0 mg or 1.0 cg, (e) 1.0 m or 1.0 yd, (f) 32 km or 20.0 mi, (g) 325 kcal or 100. J. (h) 50 nm or 0.5 m, (i) 1.0 gal or 3.5 liters (j) 80.0 km/hr or 55 mi/hr, (k) 10 lb or 10 kg, (l) 800.0 mg or 1.0 kg, (m) 0.8 nm or 8 Å?

    1.11. Place x’s in as many boxes as are appropriate for each case. Note: Brine is sea water, pencil lead is a clay-graphite mixture, gasoline contains many hydrocarbons, and Dry Ice is solid carbon dioxide.

    1.12. Express the following measurements for an ideal human body in the metric system: (a) female: 5′ 6″ tall, 125 lb in weight and 2.0 ft³ in volume and (b) male: 5′ 11″ tall, 185 lb in weight and 2.8 ft³ in volume.

    1.13. The label on a package from Europe gave the dimensions as 42 cm × 84 cm × 7 cm and the weight at 3 kg. A local postal clerk wanted to find out if postal regulations had been violated by shipping such a large package. What dimensions did the clerk arrive at after converting these metric values to the English system?

    1.14. An athlete runs 100.00 yd in 9.4 sec. If this velocity is maintained exactly, how long will it take the athlete to run 100.00 m?

    1.15. A very important constant that we will encounter in this book is known as the ideal gas constant. It is numerically equal to 8.314 J/mole °K. Express the value of this constant in (a) erg/mole °K, (b) cal/mole°K, and (c) liter atm/mole°K given 1 J = 10⁷ erg, 1 cal = 4.184 J, and 1 liter-atm = 24.2 cal.

    1.16. One hundred cubic centimeters of uranium metal weighs 1.897 kg. What is the density in lb/ft³?

    1.17. Assuming the density of water to be 1.0 g/cm³, what is the mass of a gallon of water in pounds?

    1.18. Convert the following temperatures which are commonplace in our daily lives to the Celsius scale: (a) normal body temperature, 98.6°F; (b) a cold, wintry day, 10°F; a warm fall day, 78°F; and (d) the running temperature of a modern auto engine, 250°F.

    1.19. Convert each of the following melting point temperatures to values on the Kelvin scale: (a) water, 32°F; (b) cesium, 84°F; (c) white phosphorus, 111°F; and (d) nitrogen, -346°F.

    1.20. Convert each of the following boiling point temperatures to values on the Fahrenheit scale: (a) water, 100.00X; (b) nitric oxide, -151.8°C; (c) sulfur, 444.6°C; (d) iron, 2750°C; and (e) sulfuric acid, 338°C.

    1.21*. A letter, weighed using a triple-beam balance, was 53.5 g. What would be the amount of postage required to mail the letter at these first-class rates: 15¢ for the first ounce and 13¢ for each additional ounce?

    1.22*. A dairy buys exactly 1000 qt of milk at 45¢ per quart. During the night the country changes to the metric system and the dairy sells the milk for 45¢ per liter. How much profit or loss was made?

    1.23*. The radius of a hydrogen atom is about 0.58 Å, and the distance between the sun and the Earth is about 93 million miles. Find the ratio of the radius of the hydrogen atom to the sun–Earth distance so that the units cancel.

    1.24*. The radius of a neutron (a subatomic particle) is approximately 1.5 × 10¹³ m. Find the density of a neutron if its mass is 1.675 × 10−24 g. How does this compare to the density of the heavy metal uranium, which is 19 g/cm³?

    1.25*. A container weighs 68.31 g empty and dry, 93.34 g filled with water, and 88.42 g filled with a second liquid. If the density of water is 1.0000 g/ml, find the density of the second liquid.

    1.26*. Assuming that the density of water is 1.000 g/ml, compute the density of a metal sample from the following data: (a) weight of empty container = 66.734 g, (b) weight of container and sample = 87.807 g, (c) weight of container, water and sample = 105.408 g, and (d) weight of container and water to occupy the same volume as in (c) = 91.786 g.

    1.27*. At what temperature will a Fahrenheit thermometer give (a) the same reading as a Celsius thermometer, (b) a reading that is twice that on the Celsius thermometer, and (c) a reading that is numerically the same but opposite in sign from the Celsius scale?

    1.28*. Confirm the values given in Figure 1.3 for the absolute zero shown on the Fahrenheit and Celsius scales by assuming the value of 0°K and calculating the corresponding value on each scale.

    1.29*. The British Thermal Unit, BTU, is the common unit for rating home air conditioners and furnaces. It is defined as the quantity of heat required to raise the temperature of one pound of water by one degree Fahrenheit. A calorie is the quantity of heat required to raise the temperature of one gram of water by one degree Celsius. How many calories are equivalent to one BTU?

    1.30**. A buyer of aluminum for a canning company suspected that a supply of metal cubes was not solid. A typical cube weighed 42.22 g and was 2.5 cm along an edge; (a) calculate its density. The result was compared to the value determined using a cylinder of metal known to be solid which had a radius of 2.50 cm, a length of 10.00 cm and a mass of 0.5305 kg. (b) Calculate the density of the cylindrical piece and (c) compare the results of the two densities. Note: The volume of a cube is equal to the cube of its edge, and the volume of a cylinder is equal to π(pi), or 3.1416, times the square of the radius times the length. (d) Is density a chemical or physical property? The buyer reported the density to the company’s engineering department in lb/ft³ units. (e) What was this value?

    This exercise illustrates an application of the scientific method to an everyday problem. What was the (f) hypothesis, (g) experiment, and (h) theory?

    2

    ATOMS, MOLECULES, AND IONS

    Publisher Summary

    This chapter presents the introduction of atoms, molecules, ions, and their weight relationships. It explains the behavior of the atoms. The chapter discusses the symbols used for atoms, how formulas are determined, and solutions and their concentrations. It explains three laws describing the behavior of matter that were known before the acceptance of atomic theory. The chapter further explains how atomic theory applies to those laws. Atoms are found joined together in independent particles called molecules, as ions formed by the gain or loss of electrons from the atoms and as free atoms. Ions are held together in compounds by the attraction between the positive and the negative charges of cations and anions.

    Atomic theory is presented in this chapter. Modern definitions of atoms, molecules, and ions—the three types of particles of which all substances are composed—are given. The symbols for the elements and the formulas of chemical compounds are introduced. Finally, atomic and molecular weights and the weight relationships that can be derived from chemical formulas are presented, with emphasis on Avogadro’s number and the concept of the mole and molar weight.

    How many times can you divide a piece of iron into smaller and smaller pieces that retain the properties of iron? Will you reach a point where no further division is possible, or can the process go on indefinitely? Such questions about the nature of matter were debated by the Greek philosophers. Democritus of Abdera, in about 400 b.c., came remarkably close to the modern answer to these questions—atomic theory. He argued that all matter is composed of tiny homogeneous particles which are hard and impenetrable, differ in size and shape, can come together in different combinations, and are constantly in motion. Democritus named the particles atoms from the Greek word meaning "indivisible."

    Aristotle opposed these ideas. Instead of atoms he favored the concept that Earth, air, fire, and water form the basis of all matter. Aristotle’s teachings were widely accepted and the theory of atoms fell into disrepute for over 1500 years. In the Middle Ages, progress in what was to become chemistry fell to the alchemists. Unlike the Greeks, the alchemists were experimentalists—they did not concern themselves with theories about atoms. The alchemists never reached their goals of transforming baser metals into gold, or finding the Elixir that would impart eternal life. However, from their laboratories came much practical information about metals and minerals, information that was later useful in the growth of the science of chemistry.

    Chemistry: where to begin?

    What might you do if you were about to take a trip to a foreign country? Probably you would first get a book or two from which to learn some practical things such as a few phrases in the language and the value of the money. If you planned to stay for quite a while, you could assume that these basics would help to get the trip started, but that more would have to be learned along the way. In this chapter we deal with a similar situation as we embark on the study of chemistry. There are certain things that we need to know in order to get started. Yet much must be left to be learned further down the road.

    The starting point for this journey is the atomic theory. (Some of the history of the atomic theory is given in the Aside, Toward Atomic Theory Through History.) Atoms, and the molecules and ions derived from them, are the currency of chemistry. It is not possible to go very far without understanding what they are and how to use them. The purpose of this chapter is to introduce these species and their weight relationships, while recognizing that fuller explanations of some aspects of the subject must be presented later.

    First we examine the behavior of atoms in the light of what was known when John Dalton presented his atomic theory. Then symbols used for atoms are introduced. We next define the terms molecule and ion, and extend the discussion of symbolism to these species. The very important subject of the relative weights of atoms and molecules is then presented, followed by a discussion of how formulas are determined. Finally, we briefly discuss solutions and their concentrations for the benefit of those who will soon be using solutions in the laboratory.

    2.1 What is an atom?

    We do not use the word atom in quite the same way that the ancient Greeks did. (Our word molecule is nearer to their meaning, as we will see in the next section.) However, their perception of an atom was remarkably close to what we know today. In modern terminology, the word atom is reserved for the smallest particle of an element that can participate in a chemical reaction.

    A one-sentence definition of an atom is not very informative. We can learn more about atoms by examining how atomic theory explains many observations about the behavior of matter. To begin with, atomic theory allows us to define an element as a substance composed of only one kind of atom. In Chapter 1, we defined an element as a pure substance that cannot be converted to a simpler substance by any chemical reaction. The atomic definition explains why this is so: Elements are already in the simplest form possible.

    In the following section we discuss three laws describing the behavior of matter that were known before the acceptance of atomic theory. We will see how atomic theory applies to these laws and helps to explain them.

    AN ASIDE

    Toward the atomic theory through history

    The philosophers of ancient Greece were the first to speculate about the existence of atoms—ultimate small, indivisible particles that make up all of matter. Lucretius (96–55 b.c.) said it this way, … all nature as it is in itself consists of two things—for there are atoms and there is the void.

    Over the centuries until the 1800s the theory of the existence of atoms came in and out of favor. The remarkable way in which atomic theory can explain the behavior of matter could not be fully appreciated, however, until chemistry became an experimental science. Robert Boyle (1627–1691) was a strong advocate for the experimental approach. He demonstrated how to learn by careful experimentation, and had only disdain for those who expounded theories about matter without testing their theories against facts. Boyle did not accept atomic theory nor did he believe in such elements as fire, water, salt, or sulfur, which some scientists of the time thought to be present in all matter. He did, however, show how various types of chemical substances could be distinguished from each other, and he carefully and accurately recorded his experiments so that they could be duplicated.

    For more than 100 years after Boyle, knowledge was accumulated about the behavior of minerals and other pure substances, and especially about gases. John Dalton (1766–1844), following in Boyle’s tradition of careful observation, began in 1787 to keep a journal about the weather, a habit that he continued for 57 years. A few years later Dalton began to experiment with water vapor in the air, the solubility of gases, and the combining weights of gases. It is thought that this work eventually led him to revive atomic theory.

    Dalton’s atomic theory, first published in 1808, can be summarized as follows:

    1. All matter consists of tiny particles. Dalton, like the Greeks, called these particles atoms.

    2. Atoms of one element can neither be subdivided nor changed into atoms of any other element.

    3. Atoms cannot be created or destroyed.

    4. All atoms of the same element are identical in mass, size, and other properties.

    5. Atoms of one element differ in mass and other properties from atoms of other elements.

    6. Chemical combination is the union of atoms of different elements; the atoms combine in simple, whole-number ratios to each other.

    Chemists in Dalton’s time were struggling to understand the affinity of some substances for others, which leads to their chemical combination, and the weight relationships among such substances. Joseph Gay-Lussac provided strong support for Dalton’s atomic theory in his observation that the volumes of combining gases always had simple, whole-number ratios to each other. Avogadro clearly explained the significance of this observation with the following hypothesis, now accepted as a law: Equal volumes of gas (at the same temperature and pressure) must contain equal numbers of particles. (In Chapter 3 we will further examine properties of gases.)

    Confusion was mounting during this period over the meaning of the terms atom and molecule as they were used by different groups of chemists, and also over the weights of the combining substances. Stanislao Cannizaro opened the door to a clarification of the situation when, in 1858, he brought Avogadro’s ideas, which had been largely ignored, to the attention of the scientific community. Cannizaro also proved the usefulness of Avogadro’s hypothesis by devising a method for determining the relative weights of atoms.

    Law of conservation of mass: In chemical reactions matter is neither created nor destroyed.

    2.2 Atoms and mass in chemical combination

    Three laws summarize what was known about atoms and mass in chemical combination up to the time of Dalton. These laws remain valid today for most chemical reactions and compounds.

    The first law, the law of conservation of mass, may be stated as follows: In chemical reactions, matter is neither created nor destroyed. Put another way, the total mass of the materials that react is equal to the total mass of the materials produced in the reaction. In ordinary chemical reactions it is atoms that are neither created nor destroyed. In Section 1.2 we defined a chemical reaction as a process in which at least one substance is changed in identity. This is accomplished by the rearrangement of atoms, but the atoms themselves remain unchanged.

    The law of definite proportions reflects the unvarying composition of compounds. In a pure compound, two or more elements are combined in definite proportions by weight. For example, consider sodium chloride (see Figure 2.1). We all know it as white, crystalline table salt. One way to prepare it in the laboratory is to pass chlorine, a greenish gas, over sodium, a silvery metal. The sodium burns in the chlorine, and crystalline sodium chloride appears in the container. Sodium chloride made in this way or in any other way, or sodium chloride obtained from natural sources, has been examined often enough for us to state with certainty that it is a compound containing 39.34% sodium and 60.66% chlorine by weight. Every 100.00 g of sodium chloride is the product of the combination of 39.34 g of sodium and 60.66 g of chlorine. Here is evidence that atoms of different elements have different masses. We can assume, based on the properties of sodium chloride, that each atom of sodium combines with one atom of chlorine. From the combining weights of sodium and chlorine we can then conclude that the weight of a sodium atom is roughly two-thirds that of a chlorine atom.

    FIGURE 2.1 The law of definite proportions.

    Sodium chloride NaCl is 39.34% sodium by weight and 60.66% chlorine by weight.

    The law of multiple proportions, first stated by Dalton, is a direct consequence of his atomic theory: If two elements combine to form more than one compound, a fixed weight of element A will combine with two or more different weights of element B so that the different weights of B are in the ratio of small whole numbers. This is just a way of saying that one atom of A will combine only with one atom of B (forming AB), or with two atoms of B (forming AB2), or with three atoms of B (forming AB3), and so on. For example, carbon (symbolized by

    Enjoying the preview?
    Page 1 of 1