Two or more atoms that bond together form a(n)

Chemical Relationships

by Anthony Carpi, Ph.D.

An updated version of the Chemical Bonding module is available. To view it, go here.

Though the periodic table has only 118 or so elements, there are obviously more substances in nature than 118 pure elements. This is because atoms can react with one another to form new substances called compounds (see our Chemical Reactions module). Formed when two or more atoms chemically bond together, the resulting compound is unique both chemically and physically from its parent atoms.

Let's look at an example. The element sodium is a silver-colored metal that reacts so violently with water that flames are produced when sodium gets wet.  The element chlorine is a greenish-colored gas that is so poisonous that it was used as a weapon in World War I. When chemically bonded together, these two dangerous substances form the compound sodium chloride, a compound so safe that we eat it every day - common table salt!

Two or more atoms that bond together form a(n)
Figure: The formula for table salt.

In 1916, the American chemist Gilbert N. Lewis proposed that chemical bonds are formed between atoms because electrons from the atoms interact with each other. Lewis had observed that many elements are most stable when they contain eight electrons in their valence shell. He suggested that atoms with fewer than eight valence electrons bond together to share electrons and complete their valence shells.

While some of Lewis' predictions have since been proven incorrect (he suggested that electrons occupy cube-shaped orbitals), his work established the basis of what is known today about chemical bonding. We now know that there are two main types of chemical bonding; ionic bonding and covalent bonding.

In ionic bonding, electrons are completely transferred from one atom to another. In the process of either losing or gaining negatively charged electrons, the reacting atoms form ions. The oppositely charged ions are attracted to each other by electrostatic forces, which are the basis of the ionic bond.

For example, during the reaction of sodium with chlorine:

sodium (on the left) loses its one valence electron to chlorine (on the right),
Two or more atoms that bond together form a(n)
resulting in
a positively charged sodium ion (left) and a negatively charged chlorine ion (right).
Two or more atoms that bond together form a(n)

Two or more atoms that bond together form a(n)

Interactive Animation: The reaction of sodium with chlorine

Notice that when sodium loses its one valence electron it gets smaller in size, while chlorine grows larger when it gains an additional valence electron. This is typical of the relative sizes of ions to atoms. Positive ions tend to be smaller than their parent atoms while negative ions tend to be larger than their parent. After the reaction takes place, the charged Na+ and Cl- ions are held together by electrostatic forces, thus forming an ionic bond. Ionic compounds share many features in common:

  • Ionic bonds form between metals and nonmetals.
  • In naming simple ionic compounds, the metal is always first, the nonmetal second (e.g., sodium chloride).
  • Ionic compounds dissolve easily in water and other polar solvents.
  • In solution, ionic compounds easily conduct electricity.
  • Ionic compounds tend to form crystalline solids with high melting temperatures.

This last feature, the fact that ionic compounds are solids, results from the intermolecular forces (forces between molecules) in ionic solids. If we consider a solid crystal of sodium chloride, the solid is made up of many positively charged sodium ions (pictured below as small gray spheres) and an equal number of negatively charged chlorine ions (green spheres). Due to the interaction of the charged ions, the sodium and chlorine ions are arranged in an alternating fashion as demonstrated in the schematic. Each sodium ion is attracted equally to all of its neighboring chlorine ions, and likewise for the chlorine to sodium attraction. The concept of a single molecule does not apply to ionic crystals because the solid exists as one continuous system. Ionic solids form crystals with high melting points because of the strong forces between neighboring ions.

Sodium Chloride Crystal NaCl Crystal Schematic
Two or more atoms that bond together form a(n)
Cl-1 Na+1 Cl-1 Na+1 Cl-1
Na+1 Cl-1 Na+1 Cl-1 Na+1
Cl-1 Na+1 Cl-1 Na+1 Cl-1
Na+1 Cl-1 Na+1 Cl-1 Na+1

The second major type of atomic bonding occurs when atoms share electrons. As opposed to ionic bonding in which a complete transfer of electrons occurs, covalent bonding occurs when two (or more) elements share electrons. Covalent bonding occurs because the atoms in the compound have a similar tendency for electrons (generally to gain electrons). This most commonly occurs when two nonmetals bond together. Because both of the nonmetals will want to gain electrons, the elements involved will share electrons in an effort to fill their valence shells. A good example of a covalent bond is that which occurs between two hydrogen atoms. Atoms of hydrogen (H) have one valence electron in their first electron shell. Since the capacity of this shell is two electrons, each hydrogen atom will "want" to pick up a second electron. In an effort to pick up a second electron, hydrogen atoms will react with nearby hydrogen (H) atoms to form the compound H2. Because the hydrogen compound is a combination of equally matched atoms, the atoms will share each other's single electron, forming one covalent bond. In this way, both atoms share the stability of a full valence shell.

Interactive Animation: Covalent bonding between hydrogen atoms

Unlike ionic compounds, covalent molecules exist as true molecules. Because electrons are shared in covalent molecules, no full ionic charges are formed. Thus covalent molecules are not strongly attracted to one another. As a result, covalent molecules move about freely and tend to exist as liquids or gases at room temperature.

Multiple Bonds: For every pair of electrons shared between two atoms, a single covalent bond is formed.  Some atoms can share multiple pairs of electrons, forming multiple covalent bonds. For example, oxygen (which has six valence electrons) needs two electrons to complete its valence shell. When two oxygen atoms form the compound O2, they share two pairs of electrons, forming two covalent bonds.

Lewis Dot Structures: Lewis dot structures are a shorthand to represent the valence electrons of an atom. The structures are written as the element symbol surrounded by dots that represent the valence electrons. The Lewis structures for the elements in the first two periods of the periodic table are shown below.

Two or more atoms that bond together form a(n)
Lewis Dot Structures
Two or more atoms that bond together form a(n)
Two or more atoms that bond together form a(n)
Two or more atoms that bond together form a(n)
Two or more atoms that bond together form a(n)
Two or more atoms that bond together form a(n)
Two or more atoms that bond together form a(n)
Two or more atoms that bond together form a(n)
Two or more atoms that bond together form a(n)
Two or more atoms that bond together form a(n)

Lewis structures can also be used to show bonding between atoms. The bonding electrons are placed between the atoms and can be represented by a pair of dots or a dash (each dash represents one pair of electrons, or one bond). Lewis structures for H2 and O2 are shown below.

H2 H:H or H-H
O2
Two or more atoms that bond together form a(n)
Two or more atoms that bond together form a(n)
Two or more atoms that bond together form a(n)

There are, in fact, two subtypes of covalent bonds. The H2 molecule is a good example of the first type of covalent bond, the nonpolar bond. Because both atoms in the H2 molecule have an equal attraction (or affinity) for electrons, the bonding electrons are equally shared by the two atoms, and a nonpolar covalent bond is formed. Whenever two atoms of the same element bond together, a nonpolar bond is formed.

Two or more atoms that bond together form a(n)
Figure: A water molecule: H2O.

A polar bond is formed when electrons are unequally shared between two atoms. Polar covalent bonding occurs because one atom has a stronger affinity for electrons than the other (yet not enough to pull the electrons away completely and form an ion). In a polar covalent bond, the bonding electrons will spend a greater amount of time around the atom that has the stronger affinity for electrons. A good example of a polar covalent bond is the hydrogen-oxygen bond in the water molecule.

Water molecules contain two hydrogen atoms (pictured in blue) bonded to one oxygen atom (red). Oxygen, with six valence electrons, needs two additional electrons to complete its valence shell. Each hydrogen contains one electron. Thus oxygen shares the electrons from two hydrogen atoms to complete its own valence shell, and in return shares two of its own electrons with each hydrogen, completing the H valence shells.

Interactive Animation:Polar covalent bonding simulated in water

The primary difference between the H-O bond in water and the H-H bond is the degree of electron sharing. The large oxygen atom has a stronger affinity for electrons than the small hydrogen atoms. Because oxygen has a stronger pull on the bonding electrons, it preoccupies their time, and this leads to unequal sharing and the formation of a polar covalent bond.

Because the valence electrons in the water molecule spend more time around the oxygen atom than the hydrogen atoms, the oxygen end of the molecule develops a partial negative charge (because of the negative charge on the electrons). For the same reason, the hydrogen end of the molecule develops a partial positive charge. Ions are not formed; however, the molecule develops a partial electrical charge across it called a dipole. The water dipole is represented by the arrow in the pop-up animation (above) in which the head of the arrow points toward the electron dense (negative) end of the dipole and the cross resides near the electron poor (positive) end of the molecule.

Chemical bonding between atoms results in compounds that can be very different from the parent atoms. This module, the second in a series on chemical reactions, describes how atoms gain, lose, or share electrons to form ionic or covalent bonds. The module lists features of ionic and covalent compounds. Lewis dot structures and dipoles are introduced.

Anthony Carpi, Ph.D. “Chemical Bonding (previous version)” Visionlearning Vol. CHE-4 (1), 2016.

Top


Page 2

Chemical Relationships

by Anthony Carpi, Ph.D., Adrian Dingle, B.Sc.

This is an updated version of our Chemical Bonding module. For the previous version, please go here.

Life on Earth depends on water – we need water to drink, bathe, cool ourselves off on a hot summer day (Figure 1). In fact, evidence suggests that life on Earth began in the water, more specifically in the ocean, which has a combination of water and salts, most prominently common table salt – sodium chloride. But where do water and these common salts appear on the great organizer of the elements, the periodic table? Well they, and millions of other substances, are not found on the most famous of all chemistry references: the periodic table. Why not? The answer is a simple one.

Two or more atoms that bond together form a(n)
Figure 1: Life on Earth depends on water, not only for key biological functions but also for pleasure. For example, this relaxing oasis on the Mediterranean Sea, Cala Tío Ximo beach in Benidorm, Spain. image © Diego Delso

The periodic table organizes the 118 currently recognized chemical elements, but water and sodium chloride are not elements. Rather, both are substances that are made up of a combination of elements in a fixed ratio. Such fixed ratio combinations of those 118 elements are known as compounds.

In its chemical reactions and physical interactions, sodium chloride doesn’t act like the elements that make it up (sodium and chlorine); rather, it acts as a completely different and unique substance. That’s a good thing since chlorine is a poisonous gas that has been used as a chemical weapon, and sodium is a highly reactive metal that is mildly explosive with water. So what allows sodium chloride to act in an entirely different way? The answer is that within table salt, sodium and chlorine are joined together by a chemical bond that creates a unique compound, very different from the individual elements that comprise it.

The chemical bond can be thought of as a force that holds the atoms of various elements together in such compounds. It opens up the possibility of millions and millions of combinations of the elements, and the creation of millions and millions of new compounds. In short, the existence of the chemical bonds accounts for the richness of chemistry that reaches far beyond just those 118 building blocks.

When discussing the history of chemistry it’s always dangerous to point to the specific origin of an idea, since by its very definition, the scientific process relies upon the gradual refinement of ideas that came before. However, as is the case with a number of such ideas, one can point to certain seminal moments, and in the case of chemical bonding, a famous early 18th century publication provides one such moment.

In his 1704 publication Opticks, Sir Isaac Newton makes mention of a force that points to the modern idea of the chemical bond. In Query 31 of the book, Newton describes ‘forces’ – other than those of magnetism and gravity – that allow ‘particles’ to interact.

In 1718, while translating Opticks into his native language, French chemist Étienne François Geoffroy created an Affinity Table. In this fascinating first look at the likelihood of certain interactions, Geoffroy tabulated the relative affinity that various substances had for other substances, and therefore described the strength of the interactions between those substances.

While Newton and Geoffroy’s work predated our modern understanding of elements and compounds, their work provided insight into the nature of chemical interactions. However, it was over 100 years before the concept of the combining power of elements was understood in a more modern sense. In a paper in the journal Philosophical Transactions entitled “On a new series of organic bodies containing metals” (Frankland, 1852), Edward Frankland describes the “"combining power of elements,” a concept now known as valency in chemistry. Frankland summarized his thoughts by proposing what he described as a ‘law’:

A tendency or law prevails (here), and that, no matter what the characters of the uniting atoms may be, the combining power of the attracting element, if I may be allowed the term, is always satisfied by the same number of these atoms.

Frankland’s work suggested that each element combined with only a limited number of atoms of another element, thus alluding to the concept of bonding. But it was two other scientists who performed the most important contemporary research on the concept of bonding.

In 1916, the American scientist Gilbert N. Lewis published a now famous paper on bonding entitled “The atom and the molecule” (Lewis, 1916). In that paper he outlined a number of important concepts regarding bonding that are still used today as working models of electron arrangement at the atomic level. Most significantly, Lewis developed a theory about bonding based on the number of outer shell, or valence, electrons in an atom. He suggested that a chemical bond was formed when two atoms shared a pair of electrons (later renamed a covalent bond by Irving Langmuir). His "Lewis dot diagrams" used a pair of dots to represent each shared pair of electrons that made up a covalent bond (Figure 2).

Two or more atoms that bond together form a(n)
Figure 2: Lewis dot structures for the elements in the first two periods of the periodic table. The structures are written as the element symbol surrounded by dots that represent the valence electrons.

Lewis also championed the idea of ‘octets’ (groups of eight), that a filled valence shell was crucial in understanding electronic configuration as well as the way atoms bond together. The octet had been discussed previously by chemists such as John Newland, who felt it was important, but Lewis advanced the theory.

Comprehension Checkpoint

Lewis based his theory of bonding on

While still in college, a young chemist by the name of Linus Pauling familiarized himself with Lewis’s work and began to consider how it might be interpreted within the context of the newly developed field of quantum mechanics. The theory of quantum mechanics, developed in the first half of the 20th century, had redefined our modern understanding of the atom and so any theory of bonding would be incomplete if it were not consistent with this new theory (see our modules Atomic Theory II: Bohr and the Beginnings of Quantum Theory and Atomic Theory III: Wave-Particle Duality and the Electron for more information).

Pauling’s greatest contribution to the field was his book The Nature of the Chemical Bond (Pauling, 1939). In it, he linked the physics of quantum mechanics with the chemical nature of the electron interactions that occur when chemical bonds are made. Pauling’s work concentrated on establishing that true ionic bonds and covalent bonds sit at extreme ends of a bonding spectrum, and that most chemical bonds are classified somewhere between those extremes. Pauling further developed a sliding scale of bond type governed by the electronegativity of the atoms participating in the bond.

Pauling’s immense contributions to our modern understanding of the chemical bond led to his being awarded the 1954 Nobel Prize for "research into the nature of the chemical bond and its application to the elucidation of the structure of complex substances."

Chemical bonding and interactions between atoms can be classified into a number of different types. For our purposes we will concentrate on two common types of chemical bonds, namely covalent and ionic bonding.

Molecular bonds are formed when constituent atoms come close enough together such that the outer (valence) electrons of one atom are attracted to the positive nuclear charge of its neighbor. As the independent atoms approach one another, there are both repulsive forces (between the electrons in each atom and between the nuclei of each atom), and attractive forces (between the positive nuclei and the negative valence electrons). Some constituents require the addition of energy, called the activation energy, to overcome the initial repulsive forces. But at various distances, the atoms experience different attractive and repulsive forces, ultimately finding the ideal separation distance where the electrostatic forces are reduced to a minimum. This minimum represents the most stable position, and the distance between the atoms at this point is known as the bond length.

As the name suggests, covalent bonding involves the sharing (co, meaning joint) of valence (outer shell) electrons. As described previously, the atoms involved in covalent bonding arrange themselves in order to achieve the greatest energetic stability. And the valence electrons are shared – sometimes equally, and sometimes unequally – between neighboring atoms. The simplest example of covalent bonding occurs when two hydrogen atoms come together to ultimately form a hydrogen molecule, H2 (Figure 3).

Two or more atoms that bond together form a(n)
Figure 3: Here the interaction of two gaseous hydrogen atoms is charted showing the potential energy (purple line) versus the internuclear distance of the atoms (in pm, trillionths of a meter). The observed minimum in potential energy is indicated as the bond length (r) between the atoms. image © Saylor Academy

The covalent bond in the hydrogen molecule is defined by the pair of valence electrons (one from each hydrogen atom) that are shared between the atoms, thus giving each hydrogen atom a filled valence shell. Since one shared pair of electrons represents one covalent bond, the hydrogen atoms in a hydrogen molecule are held together with what is known as a single covalent bond, and that can be represented with a single line, thus H-H.

There are many instances where more than one pair of valence electrons are shared between atoms, and in these cases multiple covalent bonds are formed. For example, when four electrons are shared (two pairs), the bond is called a double covalent bond; in the case of six electrons being shared (three pairs) the bond is called a triple covalent bond.

Common examples of such multiple bonds are those formed between atoms in oxygen and nitrogen gas. In oxygen gas (O2), two atoms share a double bond resulting in the structure O=O. In nitrogen gas (N2), a triple bond exists between two nitrogen atoms, N≡N (Figure 4).

Two or more atoms that bond together form a(n)
Figure 4: The bonds between gaseous oxygen and nitrogen atoms. In oxygen gas (O2), two atoms share a double bond resulting in the structure O=O. In nitrogen gas (N2), a triple bond exists between two nitrogen atoms, N≡N.

Double covalent bonds are shorter and stronger than comparable single covalent bonds, and in turn, triple bonds are shorter and stronger than double bonds – nitrogen gas, for example, does not react readily because it is a strongly bonded stable compound.

Comprehension Checkpoint

When four electrons are shared between atoms, _____ bonds are formed.

Ionic bonding occurs when valence electrons are shared so unequally that they spend more time in the vicinity of their new neighbor than their original nuclei. This type of bond is classically described as occurring when atoms interact with one another to either lose or gain electrons. Those atoms that have lost electrons acquire a net positive charge and are called cations, and those that have gained electrons acquire a net negative charge and are referred to as anions. The number of electrons gained or lost by a constituent atom commonly conforms with Lewis’s valence octets, or filled valence shell principle.

In reality even the most classic examples of ionic bonding, such as the sodium chloride bond, contain characteristics of covalent bonding, or sharing of electrons of outer shell electrons. A common misconception is the idea that elements tend to bond with other elements in order to achieve these octets because they are 'stable' or, even worse, 'happy', and that’s what elements 'want'. Elements have no such feelings; rather, the actual reason for bond formation should be considered in terms of the energetic stability arising from the electrostatic interaction of positively charged nuclei with negatively charged electrons.

Substances that are held together by ionic bonds (like sodium chloride) can commonly separate into true charged ions when acted upon by an external force, such as when they dissolve in water. Further, in solid form, the individual atoms are not cleanly attracted to one individual neighbor, but rather they form giant networks that are attracted to one another by the electrostatic interactions between each atom’s nucleus and neighboring valence electrons. The force of attraction between neighboring atoms gives ionic solids an extremely ordered structure known as an ionic lattice, where the oppositely charged particles line up with one another to create a rigid, strongly bonded structure (Figure 5).

Two or more atoms that bond together form a(n)
Figure 5: A sodium chloride crystal, showing the rigid, highly organized structure.

The lattice structure of ionic solids conveys certain properties common to ionic substances. These include:

  • High melting and boiling points (due to the strong nature of the ionic bonds throughout the lattice).
  • An inability to conduct electricity in solid form when the ions are held rigidly in fixed positions within the lattice structure. Ionic solids are insulators. However, ionic compounds are often capable of conducting electricity when molten or in solution when the ions are free to move.
  • An ability to dissolve in polar solvents such as water, whose partially charged nature leads to an attraction to the oppositely charged ions in the lattice.

The special properties of ionic solids are discussed in further detail in the module Properties of Solids.

Comprehension Checkpoint

Atoms that lose electrons and acquire a net positive charge and are called

Lewis used dots to represent valence electrons. Lewis dot diagrams (see Figure 1) are a quick and easy way to show the valence electron configuration of individual atoms where no bonds have yet been made.

The dot diagrams can also be used to represent the molecules that are formed when different species bond with one another. In the case of molecules, dots are placed between two atoms to depict covalent bonds, where two dots (a shared pair of electrons) denote a single covalent bond. In the case of the hydrogen molecule discussed above, the two dots in the Lewis diagram represent a single pair of shared electrons and thus a single bond (Figure 6).

Two or more atoms that bond together form a(n)
Figure 6: Two hydrogen atoms are connected by a covalent bond. This can be represented by two dots (left) or a single bar (right).

If ionic bonding and covalent bonding sit at the extreme ends of a bonding spectrum, how do we know where any particular compound sits on that spectrum? Pauling’s theory relies upon the concept of electronegativity, and it is the differences in electronegativity between the atoms that is crucial in determining where any bond might be placed on the sliding scale of bond type.

Pauling’s scale of electronegativity assigns numbers between 0 and 4 to each chemical element. The larger the number, the higher the electronegativity and the greater the attraction that element has for electrons. The difference in electronegativity between two species helps identify the bond type. Ionic bonds are those in which a large difference in electronegativity exists between two bonding species. Large differences in electronegativity usually occur when metals bond to non-metals, so bonds between them tend to be considered ionic.

When the difference in electronegativity between the atoms that make up the chemical bond is less, then sharing is considered to be the predominant interaction, and the bond is considered to be covalent. While it is by no means absolute, some consider the boundary between ionic and covalent bonding to exist when the difference in electronegativity is around 1.7 – less of a difference tends toward covalent, and a larger difference tends towards ionic. Smaller differences in electronegativity usually occur between elements that are both considered non-metals, so most compounds that are made up from two non-metal atoms are considered to be covalent.

Comprehension Checkpoint

If there is a big difference in electronegativity between two different elements, the bond between them will be

Once differences in electronegativity have been considered, and a bond has been determined as being covalent, the story is not quite over. Not all covalent bonds are created equally. The only true, perfectly covalent bond will be one where the difference in electronegativity between the two atoms within the bond is equal to zero. When this occurs, each atom has exactly the same attraction for the electrons that make up the covalent bond, and therefore the electrons are perfectly shared. This typically occurs in diatomic (two-atom) molecules such as H2, N2, O2, and those of the halogen compounds when the atoms in the bond are identical.

However, most covalent bonds occur between elements where even though the electronegativity difference is lower than 1.7, it is not zero. In these cases, the electrons are still considered shared, that is, the bond is still considered covalent, but the sharing is not perfect.

Most covalent bonds are formed between atoms of differing electronegativity, meaning that the shared electrons are attracted to one atom within the bond more than the other. As a result, the electrons tend to spend more time at one end of the bond than the other. This sets up what is known as a dipole, literally meaning ‘two poles’. One end of the bond is relatively positive (less attraction for electrons), and one end of the bond is relatively negative (more attraction for electrons). If this difference in electron affinity exists across the molecule, then the molecule is said to be polar – meaning that it will have two different, and opposite, partial charges at either end. Water (H2O) is an excellent example of a polar molecule. Electrons are not shared evenly since hydrogen and oxygen have different electronegativities. This creates dipoles in each H-O bond, and these dipoles do not cancel each other out, leaving the water molecule polar overall (Figure 7). (Read more about these bonds in our module Properties of Liquids.)

Two or more atoms that bond together form a(n)
Figure 7: In panel A, a molecule of water, H2O, is shown with uneven electron sharing resulting in a partial negative charge around the oxygen atom and partial positive charges around the hydrogen atoms. In panel B, three H2O molecules interact favorably, forming a dipole-dipole interaction between the partial charges.

When the electrons in a bond are perfectly shared, there is no dipole, and neither end of the bond carries any partial charge. When no such overall charge exists, the molecule is said to be non-polar. An example of such a non-polar molecule is hydrogen, H2. In larger molecules with multiple covalent bonds, each bond will have either no dipole or a dipole with varying degrees of partial charge. When all of these dipoles are taken into consideration in three dimensions, the uneven distribution of charge caused by the dipoles may cancel out, making the molecule non-polar.

Alternatively, there may be a partial electrical charge across the molecule, making it a polar molecule. An example of a multiple atom non-polar molecule is carbon dioxide. Electrons are not shared evenly across the C=O bonds since carbon and oxygen have different electronegativities. This creates dipoles in each C=O bond, but because these are aligned oppositely across a linear molecule, with the oxygen atoms on either side of the carbon atom, they cancel via symmetry to leave the carbon dioxide molecule non-polar (Figure 8).

Two or more atoms that bond together form a(n)
Figure 8: Electrons are not shared evenly across the C=O bonds in CO2 and thus it contains two dipoles. Since these two dipoles are opposite to one another across a linear molecule, they cancel via symmetry to leave the carbon dioxide molecule non-polar. image © Molecule: FrankRamspott/iStockphoto

We have limited our discussion to ionic and covalent bonding and the sliding scale of bond type that exists between them. However, many other types of interactions and bonds between atoms exist, notably metallic bonding (the attractions that hold metal atoms together in metallic elements), and intermolecular forces (the interactions that exist between, rather than within, covalently bonded molecules). These each involve similar electrostatic interactions to the ones described in ionic and covalent bonds, but even those extensions are far from the end of the bonding story.

In 2014, researchers found the first experimental evidence for a new type of interaction between atoms that had been predicted in the 1980s (Fleming et al., 2014). Named a "vibrational bond," the theory describes a lightweight element (in this case, an isotope of hydrogen) oscillating or "bouncing" between two much heavier atoms (in this case, bromine) and effectively holding the larger atoms together. Donald Fleming, a chemist based at the University of British Columbia in Canada, described the new bond as being "like a Ping Pong ball bouncing between two bowling balls." As research continues, we can expect to understand interactions at the molecular level with increasing sophistication, and with it, a greater understanding of what we call chemical bonding.

The millions of different chemical compounds that make up everything on Earth are composed of 118 elements that bond together in different ways. This module explores two common types of chemical bonds: covalent and ionic. The module presents chemical bonding on a sliding scale from pure covalent to pure ionic, depending on differences in the electronegativity of the bonding atoms. Highlights from three centuries of scientific inquiry into chemical bonding include Isaac Newton’s ‘forces’, Gilbert Lewis’s dot structures, and Linus Pauling’s application of the principles of quantum mechanics.

Key Concepts

  • When a force holds atoms together long enough to create a stable, independent entity, that force can be described as a chemical bond.

  • The 118 known chemical elements interact with one another via chemical bonds, to create brand new, unique compounds that have entirely different chemical and physical properties than the elements that make them up.

  • It is helpful to think of chemical bonding as being on a sliding scale, where at one extreme there is pure covalent bonding, and at the other there is pure ionic bonding. Most chemical bonds lie somewhere between those two extremes.

  • When a chemical bond is formed between two elements, the differences in the electronegativity of the atoms determine where on the sliding scale the bond falls. Large differences in electronegativity favor ionic bonds, no difference creates non-polar covalent bonds, and relatively small differences cause the formation of polar-covalent bonds.

  • HS-C4.3, HS-C6.2, HS-PS1.A3, HS-PS1.B1
  • Fleming, D.G., Manz, J., Sato, K., and Takayanagi, T. (2014). Fundamental change in the nature of chemical bonding by isotopic substitution. Angewandte Chemie International Edition, 53(50): 13706–13709.

  • Frankland, E. (1852). On a new series of organic bodies containing metals. Philosophical Transactions, 417: 417-444. Retrieved from http://rstl.royalsocietypublishing.org/content/142/417.full.pdf+html
  • Langmuir, I. (1919). The arrangement of electrons in atoms and molecules. Journal of the American Chemical Society, 41(6): 868-934.
  • Lewis, G.N. (1916). The atom and the molecule. Journal of the American Chemical Society, 38(4): 762-786.
  • Newton, I. (1704). Opticks: or, a treatise of the reflexions, refractions, inflexions and colours of light.
  • Pauling, L. (1931). The nature of the chemical bond. Application of results obtained from the quantum mechanics and from a theory of paramagnetic susceptibility to the structure of molecules. Journal of the American Chemical Society, 53(4): 1367-1400.

Anthony Carpi, Ph.D., Adrian Dingle, B.Sc. “Chemical Bonding” Visionlearning Vol. CHE-1 (7), 2003.

Top


Page 3

Reactions and Changes

by Anthony Carpi, Ph.D., Adrian Dingle, B.Sc.

(This is an updated version of the Chemical Reactions module. For the previous version, see this page.)

Chemical reactions happen absolutely everywhere. While we sometimes associate chemical reactions with the sterile environment of the test tube and the laboratory - nothing could be further from the truth. In fact, the colossal number of transformations make for a dizzying, almost incomprehensible array of new substances and energy changes that take place in our world every second of every day.

In nature, chemical reactions can be much less controlled than you’ll find in the lab, sometimes far messier, and they generally occur whether you want them to or not! Whether it be a fire raging across a forest (Figure 1), the slow process of iron rusting in the presence of oxygen and water over a period of years, or the delicate way in which fruit ripens on a tree, the process of converting one set of chemical substances (the reactants) to another set of substances (the products) is one known as a chemical reaction.

Two or more atoms that bond together form a(n)
Figure 1: A controlled fire in Alberta, Canada, set to create a barrier for future wildfires. image © Cameron Strandberg, Rocky Mountain House

Though chemical reactions have been occurring on Earth since the beginning of time, it wasn’t until the 18th century that the early chemists started to understand them. Processes like fermentation, in which sugars are chemically converted into alcohol, have been known for centuries; however, the chemical basis of the reaction was not understood. What were these transformations and how were they controlled? These questions could only be answered when the transition from alchemy to chemistry as a quantitative and experimental science took place.

Beginning in the early Middle Ages, European and Persian philosophers became fascinated with the way that some substances seemed to “transmute” (or transform) into others. Simple stones, such as those that contained sulfur, seemed to magically burn; and otherwise unimpressive minerals were transformed, like the ore cinnabar becoming an enchanting silvery liquid metal mercury when heated. Alchemists based their approach on Aristotle’s ideas that everything in the world was composed of four fundamental substances - air, earth, fire, and water (Figure 2).

Two or more atoms that bond together form a(n)
Figure 2: Aristotle believed that everything in the world was composed of four fundamental substances - air, earth, fire, and water.

As such, they proposed, and spent generations trying to prove, that less expensive metals like copper and mercury could be turned into gold. Despite their misguided approach, many early alchemists performed foundational chemical experiments - transforming one substance into another, and so it is difficult to point to a specific date or event as the birth of the idea of an ordered, quantifiable chemical reaction. However, there are some important moments in history that have helped to make sense of it.

Antoine Lavoisier was a French nobleman in the 1700s who began to experiment with different chemical reactions. At the time, chemistry still couldn’t be described as being a true, quantitative science. Most of the theories that existed to explain the way that substances changed relied upon Greek philosophy, and there was precious little experimental detail attached to the alchemist’s tinkering.

However, during the second half of the 18th century, Lavoisier performed many quantitative experiments and observed that while substances changed form during a chemical reaction, the mass of the system – or a measure of the total amount of “stuff” present – did not change. In doing so, Lavoisier championed the idea of conservation of mass during transformations (Figure 3). In other words, unlike the alchemists before him who thought that they were creating matter out of nothing, Lavoisier proposed that substances are neither created nor destroyed, but rather change form during reactions. Lavoisier’s ideas were published in the seminal work Traité élémentaire de Chimie in 1789 (Lavoisier, 1789), which is widely hailed as the birth of modern chemistry as a quantitative science.

Two or more atoms that bond together form a(n)
Figure 3: Lavoisier's Law of Mass Conservation, which states that substances are neither created nor destroyed, but rather change form during reactions. In this example, the reactants (zinc and two hydrogen chloride molecules) convert into different products (zinc chloride and dihydrogen), but no mass is lost or created.

Joseph Proust was a French actor who followed in Lavoisier’s footsteps. Proust performed dozens of chemical reactions, starting with different amounts of various materials. Over time he observed that no matter how he started a certain chemical reaction, the ratio in which the reactants were consumed was always constant. For example, he worked extensively with copper carbonate and no matter how he changed the ratio of starting reactants, the copper, carbon, and oxygen all reacted together in a constant ratio (Proust, 1804). As a result, in the last few years of the 18th century, Proust formulated the law of constant composition (also referred to as the law of definite proportions, Figure 4).

He realized that any given chemical substance (that we now define as a compound) always consisted of the same ratio by mass of its elemental parts regardless of the method of preparation. This was a huge step forward in modern chemistry since it had been previously believed that the substances formed during chemical reactions were random and disordered.

Two or more atoms that bond together form a(n)
Figure 4: An example of Proust's Law of Constant Composition, which states that any compound always consists of the same ratio by mass of its elemental parts, regardless of the method of preparation.

The English chemist John Dalton helped make sense of the laws of conservation of mass and definite proportions in 1803 by proposing that matter was made of atoms of unique substances that could not be created or destroyed (see our module Early Ideas about Matter for more information).

Dalton extended Proust’s ideas by recognizing that it was possible for two elements to form more than one compound, but that whatever the compound was, it would always contain elements combined in whole number ratios (Dalton, 1808). This observation is known as the law of multiple proportions (Figure 5) and with his atomic theory, helped to cement Lavoisier’s observations.

Two or more atoms that bond together form a(n)
Figure 5: Dalton's Law of Multiple Proportions, which states that two elements may form more than one compound, but whatever the compound was, it would always contain elements combined in whole number ratios

These advancements, taken together, laid the groundwork for our modern understanding of chemical reactions, chemical equations, and chemical stoichiometry, or the process of expressing the relative quantities of reactants and products in a chemical reaction.

Comprehension Checkpoint

____ first theorized that while substances changed form during a chemical reaction, the mass of the system did not change.

There is a staggering array of chemical reactions. Chemical reactions occur constantly within our bodies, within plants and animals, in the air that circulates around us, in the lakes and oceans that we swim in, and even in the soil where we grow crops and build our homes. In fact, there are so many chemical reactions that occur that it would be difficult, if not impossible, to understand them all. However, one method that helps us to understand them is to categorize chemical reactions into a few, general types. While not a perfect system, placing reactions together according to their similarities helps us to identify patterns, which in turn allows predictions to be made about as yet unstudied reactions. In this module, we will consider and provide some context for a few categories of reactions, specifically: synthesis, decomposition, single replacement, double replacement, REDOX (including combustion), and acid-base reactions.

No matter the type of reaction, one universal truth applies to all chemical reactions. For a process to be classified as a chemical reaction, i.e., one where a chemical change takes place, a new substance must be produced. The formation of a new substance is nearly always accompanied by an energy change, and often with some kind of physical or observable change. The physical change can be of different types, such as the formation of bubbles of a gas, a solid precipitate, or a color change. These changes are clues to the existence of a chemical reaction and are important triggers for further research by chemists.

Prior to Lavoisier’s work, it was poorly understood that there were different gases made up of different elements. Instead, various gases were commonly mischaracterized as types of "air" or air missing parts – for example, terms commonly used were "inflammable air," or "dephlogisticated air." Lavoisier thought differently and was convinced that these were different substances. He conducted experiments where he mixed inflammable air with dephlogisticated air and a spark, and he found that the substances combined to produce water. In response, he renamed inflammable air "hydrogen" from the Greek hydro for "water" and genes for "creator." In so doing, Lavoisier was identifying a synthesis reaction. In general, a synthesis reaction is one in which simpler substances combine to form another more complex one. Hydrogen and oxygen (which Lavoisier also renamed dephlogisticated air) combine in the presence of a spark to form water, summarized by the chemical equation shown below (for more on chemical equations see the section called Anatomy of a chemical equation), it represents a simple synthesis reaction.

Equation 1

2H2(g) + O2(g) → 2H2O(l)


In 1774, the scientist Joseph Priestley turned his curiosity to a mineral called cinnabar – a brick red mineral. When he placed the mineral under sunlight amplified by a powerful magnifying glass, he found that a gas was produced which he described as having an “exalted nature” because a candle burned in the gas brightly (Priestley, 1775). Without realizing it, Priestley had discovered oxygen as a result of a decomposition reaction. Decomposition reactions are often thought of as the opposite of synthesis reactions since they involve a compound being broken down into simpler compounds or even elements. In the case of Priestley’s oxygen, he had broken down mercury (II) oxide (cinnabar) with heat into its individual elements. The reaction can be summarized in the following equation.

Equation 2

2HgO(s) → 2Hg(l) + O2(g)


The British chemist and meteorologist John Daniell, invented one of the very first practical batteries in 1836 (Figure 6). In his cell, Daniell utilized a very common single replacement reaction. His early cells were complicated affairs, with ungainly parts and complicated constructs, but by contrast, the chemistry behind them was really quite simple.

Two or more atoms that bond together form a(n)
Figure 6: Daniell cell batteries.

In certain chemical reactions, a single constituent can substitute for another one already joined in a chemical compound. The Daniell cell works because zinc can substitute for copper in a solution of copper sulfate, and in so doing exchange electrons that are used in the battery cell. The reaction can be summarized as follows:

Equation 3

Zn(s) + CuSO4(aq) → ZnSO4(aq) + Cu(s)


This particular single displacement is called a metal displacement since it involves one metal replacing another metal, and many types of batteries are based on metal replacement reactions. However, several other types of single replacement reactions exist, such as when a metal can replace hydrogen from an acid or from water, or a halogen can replace another halogen in certain salt compounds.

The controlled use of fire was a crucial development for early civilization. While it’s difficult to pin down the exact time that humans first tamed the combustion reactions that produce fire, recent research suggests it may have occurred at least a million years ago in a South African cave (Berna et al. 2012).

Chemically, combustion is no more than the reaction of a fuel (wood, oil, gasoline, etc.) with oxygen. For combustion to take place there must be a fuel and oxygen gas. However, these reactions often require activation energy (discussed in more detail in the module Chemical Bonding: The Nature of the Chemical Bond), which can be provided by a ‘spark’ or source of energy for ignition. Fuel, oxygen, and energy are the three things make up what is known as the fire triangle (Figure 7), and any one of them being absent means that combustion will not take place.

Two or more atoms that bond together form a(n)
Figure 7: The fire triangle is made up of three things - fuel, oxygen, and energy. image © Gustavb

In the modern world, many of the fuels that are typically burned for energy, are hydrocarbons – substances that contain both hydrogen and carbon (as discussed in more detail in our Carbon Chemistry module). Plants produce hydrocarbons when they grow, and thus make an excellent fuel source, and other hydrocarbons are produced when plants or animals decay over time (such as natural gas, oil, and other substances). When these fuels combust, the hydrogen and carbon within them combine with oxygen to produce two very familiar compounds, water, and carbon dioxide. One simple example is the combustion of natural gas, or methane, CH4:

Equation 4

CH4(g) + 2O2(g) → CO2(g) + 2H2O(l)


As with the combustion of all fuels, heat and light are products, too, and it is these products that are used to cook our food or to heat our homes.

Each of the four types of reaction above are sub-categories of a single type of chemical reaction known as redox reactions. A redox reaction is one where reduction and oxidation take place together, hence the name. The individual processes of oxidation and reduction can be defined in more than one way, but whatever the definition, the two processes are symbiotic, i.e., they must take place together.

In one definition, oxidation is described as the process in which a species loses electrons, and reduction is a process where a species gains electrons. In this way, we can see how the pair must take place together. If a chemical substance is to lose electrons (and therefore be oxidized), then it must have another, interdependent chemical substance that it can give those electrons to. In the process, the second substance (the one gaining electrons) is said to be reduced. Without such an electron acceptor, the original species can never lose the electrons and no oxidation can take place. When the electron acceptor is present, it gets reduced and the redox combination process is complete. Redox reactions of this type can be summarized by a pair of equations – one to show the loss of electrons (the oxidation), and the other to show the gain of electrons (the reduction). Using the example of the Daniell cell above,

Equation 5

Oxidation: Zn → Zn2+ + 2e-

Reduction: Cu2+ + 2e- → Cu


The electrons shown being lost by zinc in the first reaction, are the same electrons being accepted by the copper ions in the second. Together, the reactions can be combined to cancel out the electrons on either side of the reactions, into the overall redox reaction:

Equation 6

Zn + Cu2+ → Zn2+ + Cu


Other definitions of oxidation and reduction also exist, but in every case, the two halves of the redox reaction remain symbiotic – one loses and the other gains. The loss from one species cannot happen without the other species gaining.

When soap won’t easily produce a lather in water, the water is said to be ‘hard’. Hard water causes all kinds of problems that go beyond just making it difficult to form a lather. The buildup of compounds in water pipes (known as ‘scale’), can block the flow of water and can cause problems in industrial processes. Textile manufacturing and the beverage industry rely heavily on water. In those situations, the quality of the water can make a difference to the end product, so controlling the water composition is crucial.

Hard water contains magnesium or calcium ions in the form of a dissolved salt such as magnesium chloride or calcium chloride. When soap (sodium stearate) comes into contact with either of those salts, it enters into a double displacement reaction that forms the insoluble precipitate known as ‘soap scum’.

A double displacement reaction (also known as a double replacement reaction) occurs when two ionic substances come together and both substances swap partners. In general:

Equation 7

AB + CD → AD + CB


Where A and C are cations (positively charged ions), and B and D are anions (negatively charged).

In the case of the reaction of soap with calcium chloride, the reaction is:

Equation 8

CaCl2(aq) + 2NA(C17H35COO)(aq) → 2NaCl(aq) + Ca(C17H35COO)2(s)


The solid calcium stearate is what we call soap scum, which is formed by the reaction of the soluble sodium stearate salt (the soap) in a double replacement reaction with calcium chloride.

Acid-base reactions happen around, and even inside of us, all the time. From the classic elementary school baking soda volcano to the process of digestion, we encounter acids and bases on a daily basis.

When a hydrogen atom loses its only electron, it forms a positive ion, H+. This hydrogen ion is the essential component of all acids, and indeed one definition of an acid is that of a hydrogen ion donor. Compounds such as the citric acid in lemon juice, the ethanoic acid in vinegar, or a typical laboratory acid like hydrochloric acid, all give their hydrogen ions away in chemical reactions known as acid-base reactions. The chemical opposites of acids are known as bases, and bases can be defined as hydrogen ion acceptors. Whenever an acid donates a hydrogen ion to a base, an acid-base reaction has taken place, for example, when hydrochloric acid donates a hydrogen ion to a base such as sodium hydroxide:

Equation 9a

HCl(aq) + NaOH(aq) → H2O(l) + NaCl(aq)


A closer look at this reaction reveals that in water the HCl gives away an H+ as shown below:

Equation 9b

HCl(aq) + H2O(l) → H3O+(aq) + Cl-(aq)


The resulting species, H3O+ (the hydronium ion), can, in turn, act as an acid when it comes into contact with any species that can accept a hydrogen ion, such as hydroxide ions from sodium hydroxide:

Equation 9c

H3O+(aq) + NaOH(aq) → 2H2O(l) + Na+(aq)


Combining equations #9a and #9b gives us equation #9c.

Equation #9c can be re-written to show the individual ions that are found in solution, thus:

Equation 9d

H+(aq) + Cl-(aq) + Na+(aq) + OH-(aq) → H2O(l) + Na+(aq) + Cl-(aq)


Removing the spectator ions from the equation above, we get the net ionic equation:

Equation 9e

H+(aq) + OH-(aq) → H2O(l)


Any chemical reaction that forms water from the reaction between an acid and base as in equation #9e is known as a neutralization reaction.

Comprehension Checkpoint

The type of chemical reaction where a single constituent can substitute for another one already joined in a chemical compound is:

Chemical equations are always linked to chemical reactions since they are the shorthand by which chemical reactions are described. That fact alone makes equations incredibly important, but equations also have a crucial role to play in describing the quantitative aspect of chemistry, something that we formally call stoichiometry.

All chemical reactions take on the same, basic format. The starting substances, or reactants, are listed using their chemical formula to the left-hand side of an arrow, with multiple reactants separated with plus signs. In the case of a reaction between carbon and oxygen:


To the right hand of the arrow one finds the chemical formulas of the new substance or substances (known as the products) that are produced by the chemical reaction. In this case, since carbon dioxide is the result of burning carbon in the presence of oxygen:

Equation 10b

[Reactants] C + O2 → CO2 [Products]


Since reactions can result in both physical as well as chemical changes, each substance is given a state symbol written as a subscript to the right of the formula, this describes the physical form of the reactants and products. Common state abbreviations are (s) for solids, (l) for liquids, (g) for gases and (aq) for any aqueous substances, i.e., those dissolved in water.

Equation 10c

C(s) + O2(g) → CO2(g)


Finally, in order to ensure that this representation abides by the law of conservation of mass, the equation may need to be balanced by the addition of numbers in front of each species that create equal numbers of atoms of each element on each side of the equation. In the case of the formation of carbon dioxide from carbon and oxygen, there is no need for the addition of such numbers (called the stoichiometric coefficients), since 1 carbon atom and 2 oxygen atoms appear on each side of the equation.

In nature, chemical reactions are often driven by exchanges in energy. In this respect, reactions are generally separated into two categories – those that release energy and those that absorb energy.

Exothermic reactions are those that release energy to the surroundings (Figure 8, right). Combustion reactions are an obvious example because the energy released by the reaction is converted into the light and heat seen in the immediate surroundings.

By contrast, endothermic reactions are those that absorb energy from the surroundings (Figure 8, left). In this situation, one may have to heat up the reaction or add some other form of energy to the system before seeing the reaction proceed.

Two or more atoms that bond together form a(n)
Figure 8: On the left is an endothermic reaction, where energy is absorbed from the surroundings. In contrast, on the right is an exothermic reaction, which releases energy into the surroundings.

In both cases it is important to note that energy is neither created nor destroyed, rather it is transferred from one type of energy to another, for example from chemical energy to that of heat or light. The energy that goes into the formation of chemical bonds is exchanged for other types of energy with the environment around that reaction. A classic example is the photosynthesis reaction, in which plants absorb light energy from the sun in order to create bonds between atoms that make up sugars, which are stored as chemical energy for later use by the plant. The process of respiration is essentially the reverse of photosynthesis, where the bonds in sugar molecules are broken and the released energy is then used by the plant.

Comprehension Checkpoint

_____ reactions are those that absorb energy from the surroundings.

Chemical reactions happen all around us every day. Whether it is a single replacement reaction in the battery of our flashlight, a synthesis reaction that occurs when iron rusts in the presence of water and oxygen, or an acid-base reaction that happens when we eat – we experience chemical reactions in almost everything we do. Understanding these reactions is not an abstract concept for a chemist in a far off laboratory, rather it is critical to understanding life and the world around us. To truly master chemical reactions, we need to understand the quantitative aspect of these reactions, something referred to as stoichiometry, and a concept we will discuss in another module.

This modules explores the variety of chemical reactions by grouping them into general types. We look at synthesis, decomposition, single replacement, double replacement, REDOX (including combustion), and acid-base reactions, with examples of each.

Key Concepts

  • The steps from a qualitative science to quantitative one, were crucial in understanding chemistry and chemical reactions more completely.
  • When a substance or substances (the reactants), undergo a change that results in the formation of a new substance or substances (the products), then a chemical reaction is said to have taken place.
  • Mass and energy are conserved in chemical reactions. Matter is neither created or destroyed, rather it is conserved but rearranged to create new substances. No energy is created or destroyed, it is conserved but often converted to a different form.
  • Chemical reactions can be classified into different types depending on their nature. Each type has its own defining characteristics in terms of reactants and products.
  • Chemical reactions are often accompanied by observable changes such as energy changes, color changes, the release of gas or the formation of a solid.
  • Energy plays a crucial role in chemical reactions. When energy is released into the surroundings the reaction is said to be exothermic; when energy is absorbed from the surroundings the reaction is said to be endothermic

  • HS-C5.4, HS-PS1.A2, HS-PS1.A3, HS-PS1.B3
  • Berna, F., Goldberg, P., Horwitz, L. K., Brink, J., Holt, S., Bamford, M., & Chazan, M. (2012). Microstratigraphic evidence of in situ fire in the Acheulean strata of Wonderwerk Cave, Northern Cape province, South Africa. Proceedings of the National Academy of Sciences, 109(20), E1215-E1220.

  • Dalton, John (1808). A New System of Chemical Philosophy.
  • Lavoisier, Antoine (1789). Traité Élémentaire de Chimie, présenté dans un ordre nouveau, et d'après des découvertes modernes.
  • Priestley, Joseph (1775). "An Account of Further Discoveries in Air". Philosophical Transactions. 65: 384–94.
  • Proust Joseph Louis (1804). “Sur les Oxydations Métalliques.” J Phys. 59: 321-343.

Anthony Carpi, Ph.D., Adrian Dingle, B.Sc. “Chemical Reactions” Visionlearning Vol. CHE-1 (6), 2003.

Top


Page 4

On July 26, 1943, a black cloud enveloped much of Los Angeles. It was difficult to see more than a few blocks and people’s eyes and noses burned. In the midst of World War II, many residents feared they were under chemical attack. 

Fortunately, the dark fog was not the result of chemical warfare, but residents couldn’t identify the culprit. At first, many blamed pollution from a manufacturing plant that made synthetic rubber. Public pressure to stop the toxic cloud forced the plant to close, but the problem persisted.

People began calling the periodic gaseous invasions “smog,” thinking (incorrectly) that it was a combination of smoke and fog. Scientists thought (correctly) that emissions from cars and manufacturing plants could cause smog, but they didn’t understand how.

What is is smog? Where does it come from? And how could the city of Los Angeles stop it?

Two scientists, Arie Jan Haagen-Smit and Eugene Houdry, were instrumental in beginning to answer these questions. To do so, they used their knowledge of chemical reaction kinetics, which is the focus of this learning module.

In a chemical reaction, the speed that reactant molecules convert into product molecules is called its reaction rate. The study of reaction rates and what variables can speed them up or slow them down is called chemical reaction kinetics. 

Chemical reactions involve changes in energy (see “Energy Changes” in Chemical Reactions), and each reaction has an energy barrier, called activation energy, that reactants must overcome before they can form products.

Imagine riding a bike over a steep hill. Your legs push hard as the incline gets steeper and steeper. Finally, after spending much energy, you reach the top and—at last—coast down the other side. This is a lot like what happens with chemical reactions. Activation energy is the hill that reactants must climb over to form products. 

Cars are culprits in the Los Angeles smog mystery because of the reactions that occur during and after combustion. To understand this, let’s consider how a car engine burns gasoline to produce power. Although gasoline contains many different hydrocarbon molecules—for now—assume it is made primarily of octane (C8H18) and follows the combustion reaction in Eq. 1 (see “Combustion reactions” in Chemical Reactions):

Equation 1

2 C8H18(l) + 25 O2(g) 16 CO2(g) + 18 H2O(g)

At first glance, the reaction looks pretty clean, producing just CO2 (which is a problem for global warming but does not produce smog) and water. And an energy diagram for the reaction would look like Figure 1, where the height of the hill represents the activation energy. 

Two or more atoms that bond together form a(n)
Figure 1: Reactants must climb the activation energy hill in order to form products.

Despite what Equation 1 shows, what would happen if you simply mixed air and gasoline together? Would it spontaneously combust? No. You must first provide enough energy (the activation energy) to get the reaction going. Cars have spark plugs that shoot tiny electrical sparks to add the energy to initiate the reaction. Then, because combustion is exothermic (see “Energy Changes” in Chemical Reactions), it generates heat that sustains the reaction.

Comprehension Checkpoint

Activation energy is ______.

Just like in our bicycle example, there are ways to speed up chemical reactions. What if a friend of yours was running behind you and pushing you up the hill? You would have more energy so it’s able to climb the hill much faster. The same thing occurs in chemical reactions. Increasing the energy of reactants speeds reaction up, and decreasing the energy of reactants slows reaction rates down. But why?

When reactant molecules collide with enough energy—energy greater than or equal to the activation energy—they form products. This is known as the collision theory of reactions. If we add energy to reactant molecules, more molecules collide with enough force to push them over the activation energy hill. In other words, more reactant molecules cross the activation energy threshold in a given time, and the reaction rate increases. If we decrease the energy of reactants, the opposite occurs. 

Equation 1 suggests that gasoline combustion produces only carbon dioxide and water. But these products don’t contribute to Los Angeles’ smog problem. Arie Jan Haagen-Smit, a biochemist who had previously studied the compounds responsible for the flavor of pineapple and some wines, suspected there was more to the story.

Haagen-Smit became interested in the Los Angeles smog issue because of the damage that the pollution was doing to crops in the area. He knew that car engines did not only give off carbon dioxide and water, and that the reactions were more complicated.

It turns out that gasoline contains small amounts of sulfur compounds. When the sulfur compounds are provided with activation energy, as they are in an internal combustion engine, they react with oxygen to form sulfur oxides. Since many different compounds of sulfur and oxygen molecules are formed, including SO, SO2, etc., sulfur oxides are written as SOX. We also know that air contains nitrogen. Similar to sulfur, nitrogen reacts with oxygen during combustion to form nitrogen oxides (written as NOX for the same reason as SOX).

In addition, inefficient combustion of the carbon in gasoline forms carbon monoxide (CO) along with carbon dioxide (CO2). Since engines are not 100% efficient, incomplete combustion will leave some organic compounds in the fuel unburned (for example methane, CH4). These are usually referred to as volatile organic carbons (VOCs).

So, if we revisit our combustion reaction, we see that it is not as simple as Equation 1. Multiple reactions are actually occurring at the same time:

Equations 2A through 2C

A. Gasoline(l) + O2(g) → CO(g) + CO2(g) + H20(g) + VOCs(g)

B. N2 + O2(g) → NOx(g)

C. S + O2(g) → SOx(g)

In the 1940s, the products on the right side of Equations 2A through 2C all exited car tailpipes. Haagen-Smit suspected emissions from cars were at least partially to blame for Los Angeles’s smog problem. To prove his theory, he needed to find out what compounds were causing damage to the area’s crops.

He began to investigate smog’s chemical composition, cooling samples of the gas until some of its ingredients condensed. Then he analyzed the resulting liquid’s chemical properties. Haagen-Smit found it contained the same kind of volatile organic compounds (VOC) that could result from incompletely burned gasoline.

Now he needed to find out if these VOCs resulted in smog’s toxic properties. Working with a colleague, he exposed plants like spinach, beets, and alfalfa to smoggy air. Haagen-Smit noticed that the toxic air turned some plant leaves silver, others bronze, and others a bleachy white. Then he exposed the plants to various VOCs, thinking that the culprit chemical would produce the same discoloration of plant leaves as the smog. But after testing 50 different VOCs, the experiments failed—none of the compounds damaged the plants in the same way, and he was no closer to the answer.

Finally, he tested the plants by exposing them to a mix of VOCs and the oxidant ozone (O3). 

At last! When he exposed the plants to the mixture of ozone and VOCs, their discoloration matched that of the smog. He deduced that that difference was the ozone, and he was confident he’d found the culprit. However, he still didn’t know how ozone formed in air near the ground. 

High up in the atmosphere, the ozone layer blocks the most dangerous form of UV radiation from reaching Earth’s surface (see “Temperature in the atmosphere” in Composition of Earth’s Atmosphere). Way up there, ozone is good. Near the ground, where humans and other animals can breathe it, ozone is bad because it causes respiratory health issues like asthma and lung infections. An example of how smog can change the air quality of an area is shown in Figure 2.

Two or more atoms that bond together form a(n)
Figure 2: The word “smog” originated as a combination of smoke and fog. Over time, scientists showed that smog is actually due to ground-level ozone plus a mix of other pollutants. This is shown on the left side of this image taken of Fahne, China. The right side is 10 days later, when weather conditions had cleared the smog. image © CC BY-SA 4.0 Tomskyhaha

Haagen-Smit remembered reading that compounds could oxidize in air in the presence of sunlight, and he developed an experiment that synthesized ozone using VOCs, nitrogen oxides (NOx), and sunlight. Making smog in the lab proved it was possible for harmful ozone to form at ground level from the chemicals in automobile exhaust under the right conditions. Today, we understand ground-level ozone pollution forms according to the following reaction:

Equation 3

NOx(g) + VOC(g) + Heat + Sunlight → O3(g)

At cool temperatures and low levels of sunlight (like in winter or northern latitudes), this reaction would be so slow that we wouldn’t need to worry about ground-level ozone pollution forming. So how do heat and sunlight speed this reaction up? They add energy! Not only do they add energy to push a few reactants over the activation energy hill, they add so much energy that reaction speeds increase enough to cause an ozone pollution problem. 

Besides heat and sunlight, can you think of any other ways to increase or decrease the energy of reactants (and therefore speed up or slow down reaction rates)? Well, read on.

Comprehension Checkpoint

True or False: Volatile organic compounds (VOCs) alone cause discoloration in crops and are responsible for smog.

Collision theory tells us that heat gives reactants more kinetic energy, which causes them to move around faster (like how boiling water moves around the pot). When reactant molecules move faster, more collide with enough energy to react, and the reaction rate speeds up. Colder temperatures do the opposite. 

All of the car emission pollutant reactions shown in Equations 2A through 2C occur faster because gasoline combustion is exothermic and gives off heat. If the temperature of combustion could be reduced in a way that still allowed the reaction to power a car, these pollution reactions might not be a problem. So far, no one has discovered a way to do this.

Concentration is the amount of a substance in a given volume of solution. The higher the concentration of a substance, the more molecules there are within the same volume. A lower concentration has fewer molecules. For example, adding sugar to a hot cup of coffee or tea would increase the concentration of sugar with every spoonful. 

In areas with more manufacturing plants and traffic, concentrations of NOx and VOCs in the atmosphere are higher. These elevated concentrations mean more ozone-producing reactions, which explains why highly populated urban areas suffer more from ground-level ozone pollution than rural locations.

A cube of sugar in coffee or tea might take a few minutes to completely dissolve, whereas, a spoonful of granulated sugar would dissolve almost instantaneously. The difference comes down to surface area. In the cube, most of the sugar molecules aren’t exposed to the liquid until the outer layers dissolve. With the granulated sugar, however, the liquid tea or coffee immediately surrounds each individual grain.

It works similarly in chemical reactions that involve solids. Increasing the surface area of the solid reactants by grinding them into a fine powder exposes more molecules, increasing the likelihood that they’ll collide. 

If you wanted a cube of sugar to dissolve faster in a cup of coffee or tea, what would you do? Grab a spoon and stir it! Mixing reactants together is another way to add kinetic energy: Increase the number of reactant molecules that collide with one another and you will increase reaction rates.

Sunlight, temperature, concentration, surface area, and mixing aren’t the only way to affect reaction rates. Now that Haagen-Smit had found that car and manufacturing emissions indirectly produced smog via reactions that occurred in the atmosphere, a French mechanical engineer named Eugene Houdry would use another method of controlling reaction rates to begin solving the problem of ozone pollution.

As smog continued to be covered in the news, Eugene Houdry also became concerned about smog pollution in Los Angeles and other large cities. Houdry was especially worried because the number of automobiles and industrial plants was rapidly growing at the time. He knew the problem would only get worse.

Haagen-Smit’s work had identified the culprits in the formation of smog. Houdry began building on this work by developing a device to reduce the VOC emissions in automobile tailpipes. Houdry studied catalysts that could speed up the rates of chemical reactions that would reduce VOC pollution from forming. He used a material that lowers activation energy, called a catalyst.

Catalysts speed up reactions by lowering the activation energy required (Figure 3). Unlike the variables listed above, catalysts do not increase the energy of reactants or the number of collisions, and catalysts are not chemically changed as part of the reaction. They just make the energy barrier smaller. In the analogy of the bike riding over a hill, a catalyst shrinks the hill, making it easier to pedal to the other side (Figure 3). 

Two or more atoms that bond together form a(n)
Figure 3: Catalysts speed up reactions by lowering the activation energy.

Ironically, Houdry first began his work on catalysts in the 1920s on a chemical process used to make gasoline. The process he developed increased the amount of gasoline that could be extracted from oil and other petroleum products and helped make the growth of automobiles possible. However, Houdry had no idea then that gasoline emissions would create ozone pollution. Decades later, in the 1950s, Houdry used his knowledge of kinetics and catalysts to invent the “catalytic converter” to reduce smog emissions. Houdry’s first catalytic converter helped break down the VOCs and carbon monoxide that could be released by cars. Since VOCs were a part of the reaction that formed ozone, they also reduced smog in the process.

Comprehension Checkpoint

Catalysts reduce the emission of VOCs from cars by ________.

Over time, scientists and engineers have improved on Houdry’s design. Today, converters are more efficient at stopping pollution by reducing more pollutants cars emit. However, they haven’t completely solved the problem of ozone pollution because catalytic converters are not yet perfectly efficient, and cars are not the only source of these pollutants. It’s true that cars today emit significantly fewer pollutants than in the 1950s, but there are also many more cars on the road, and all that pollution adds up, especially in larger cities with high traffic.

Ground-level ozone is just one of many problems that reaction kinetics and catalysts can help solve. What about all the CO2 released from car tailpipes and other sources like power plants that burn coal and natural gas? Well, scientists are studying catalysts that can break down CO2 into carbon and oxygen gas. 

Earth’s climate is changing. Hopefully advances in science and alternative fuels will help us tackle climate change as well as ozone pollution. Some scientists have developed new catalysts that make jet fuel from carbon dioxide. Imagine a world where instead of drilling for more oil and gas, we simply made fuel from CO2 harvested from the atmosphere. 

This module explores the study of reaction rates and the variables that can speed them up or slow them down, also known as chemical reaction kinetics. Through real-life examples, we examine our understanding of chemical reaction products, and the developments scientists have made in manipulating them.

Key Concepts

  • Current scientific theory suggests that chemical reactions are driven by collisions between molecules - and factors that affect those collisions affect chemical reactions.

  • The study of chemical reaction rates, and the variables affecting those rates, is referred to as Chemical Reaction Kinetics. And the rate of a chemical reaction is the speed at which reactant molecules convert into product molecules.

  • By collecting data about reaction rates during experiments, scientists have recognized that most reactions do not happen spontaneously, but rather need a certain amount of energy to get them started. This minimum energy needed to “activate” the reactants is called a reaction’s activation energy.

  • Because reactions are driven by collisions, and the speed at which reactants are moving affect those collisions, chemical reactions are driven by changes in energy, and changing the energy of the reactants changes the rate of a chemical reaction.

  • Experimentation conducted by scientists have shown that several factors influence reaction rates, including temperature, surface area, concentration, pressure, mixing, and sunlight (for photochemical reactions only).

  • Many scientists have observed that certain substances can help speed up chemical reactions although they are not chemically changed in the process. These substances are called catalysts, and they work by reducing a chemical reaction’s activation energy.

Andy Carstens, MA “Chemical Reactions II” Visionlearning Vol. CHE-5 (2), 2022.


Page 5

Chemical Relationships

by Robin Marks, M.A., Anthony Carpi, Ph.D.

Imagine that you’re traveling in an exotic place and your rusty car muffler falls off. You need to find a place where you can buy a replacement, and you don’t speak the local language. Take a moment now to draw a sketch that conveys that you’re searching for a shop where you can get a new part installed. Maybe you include a car, a muffler, a storefront, and a person holding a screwdriver or other tool.

Then look at your picture: think about what kind of understanding you and the viewer need to share in order for you to convey your message. If you showed the same picture to someone 500 years ago, they’d have no idea what a car is, let alone what a muffler does or that there are people who specialize in installing them. But you and your 21st-century viewer can translate your sketch because of shared knowledge you have about cars.

Chemical equations play a similar role for people conveying messages about what happens during a chemical reaction. You probably remember that In a chemical reaction, bonds between atoms in a compound are broken, and the atoms rearrange to form new compounds, either releasing or consuming energy in the process (see our Chemical Reactions module).

As an example, let’s consider what happened to the muffler. Simply saying it rusted isn’t much of an explanation. You could say iron reacted with oxygen to produce rust. That's better, but not very precise. What, exactly, is rust? Chemically, it’s iron oxide, but iron forms many types of oxides. So we need a very specific way to express the chemical reaction that to led to our muffler’s demise.

That’s what chemical equations are for: they’re a type of shorthand used to precisely communicate exactly what’s happening in the reaction. In the most basic sense, a chemical equation describes the type and amount of each substance that reacts (the reactants) to form a given amount of specific substances produced (the products). Reactants will always react in proportions given in the equation. If the supply of one reactant runs out, the excess of the other will remain unreacted.

The shorthand that explains our rusty car part is this:

The large number in front of an atom or molecule, which we call the coefficient, tells us the relative amounts of each substance involved in or produced by the reaction. The numbers in subscript refer specifically to the element that’s in front of them.

In other words, 4 iron atoms in the muffler reacted with 3 oxygen molecules in the air. Each of these oxygen molecules contains 2 oxygen atoms. When the chemical bonds reassemble and these reactants combine, the result is 2 molecules of a specific iron oxide that contains 2 iron atoms and 3 oxygen atoms—a.k.a rust.

Two or more atoms that bond together form a(n)
Plenty of the iron in this muffler combined with oxygen in the environment, creating rust and leaving holes in the muffler where the iron has been consumed. image © Raymond Webber

This equation conveys something much deeper than numbers of particles, though: it captures the centuries-long accumulation of knowledge about what our universe is made of and how matter interacts. Like an elegant poem (or like your rusty muffler drawing), a chemical equation conveys a world of complex concepts in just a few expressions.

Comprehension Checkpoint

If one reactant is completely consumed, the remaining reactant will

As our understanding of chemical processes deepened over time, chemical equations slowly became more sophisticated. Ultimately, such equations played a role in the recognition of chemistry as its own important science, separate from medicine, alchemy (which was popular in the 17th and 18th centuries), and physics.

The first known written chemical equation—which is really more of a diagram—appeared in what’s considered to be the first chemistry textbook, the Tyrocinium Chymicum (meaning “Begin Chemistry”). It was penned by French scientist and teacher Jean Beguin in Paris in 1615, and describes what Beguin observed when he heated antimony sulfide with a chloride of mercury. The mercury became vapor, leaving behind a residue of antimony oxychloride.

While Beguin’s diagram looks very different than modern chemical equations (and it isn’t entirely correct), it reflects an understanding of reactants and products in a chemical reaction.

Two or more atoms that bond together form a(n)
Figure 2: This drawing, from Jean Beguin’s Tyrocinium Chymicum, shows what the author believed was happening when antimony sulfide reacts with mercury chloride. The diagram conveys an understanding that a chemical reaction begins with specific reactants and produces specific products.

Beguin revealed the beginnings of an understanding of what occurs in a chemical reaction, but not an explanation of why. More than a century passed before new diagrams revealed deeper insights into the phenomena that drive reactions.

These insights came from William Cullen, who was also a teacher, and founded the chemistry department at the University of Glasgow in 1747. His hand-written lecture notes contain diagrams using arrows and letters, indicating four different types of reactions:

Two or more atoms that bond together form a(n)
Figure 3: Drawings in a set of undated lecture notes written by William Cullen in the mid-18th century. Cullen developed the drawings in the hope that they would help his students better understand chemical reactions. image © University of Glasgow Library

The diagram on the lower left described what Cullen called a “single elective attraction,” or what we now call a single replacement reaction, one that involves a single element taking the place of another element within a compound. (For more on reaction types, see our Chemical Equations module. The upper left is his diagram for a double replacement reaction, or one that involves the exchange of a component from each of two different compounds. In both cases, the reactions produce precipitates, indicated by the squiggly lines.

The two diagrams on the right relate to the dissolution of salts. The prevailing wisdom during Cullen’s time was that “elective attractions,” which we now know as charge, caused a metal in one substance to be attracted to another substance. The metals swapped places, forming either a new solid precipitate or dissolving in solution (sometimes producing a solution that displayed new properties).

While he didn’t have all the details right, Cullen’s thinking was on track, and his diagrams represent an important step toward what is now a common generic chemical equation:

We may take this simple type of equation for granted now, but at the time, it demonstrated some very perceptive ideas about what is actually happening during a chemical reaction. Remember, atoms and molecules were still not understood in Cullen’s time. It took a trio of chemists working over the next 50 years to clarify the picture that gave us the modern chemical equation. So, to grasp the idea that substances combined and were exchanged in chemical reactions is really significant.

In 1774, French chemist Antoine Lavoisier made an important observation – he noted that while substances in a chemical reaction changed form in ways described by Cullen, the mass of the system did not change. In other words, the amount of each element present remained the same, meaning that matter and mass are conserved. This is a crucial concept that will be important when we discuss balancing chemical reactions – all elements have to be fully accounted for at the start and at the end of a chemical reaction.

Around the same time as Lavosier, Joseph Proust, another Frenchman, was working extensively with copper carbonate. He found that regardless of how he changed the ratio of starting reactants – adding more copper sometimes, or more carbon or oxygen at others, the copper, carbon, and oxygen all reacted together in a constant ratio. His insight brought us the law of definite proportions: in any given compound, the elements occur in fixed ratios, regardless of their source. (Visit our module on Chemical Reactions to learn more about the work of these scientists). Again, this seems obvious to us today as we know that elements only react together in certain ways, but Proust completed his work before it was widely recognized that atoms and compounds existed.

Finally, in 1803, English chemist John Dalton tied these threads together by proposing that matter is made of atoms of unique substances that could not be created or destroyed (see our module Early Ideas about Matter for more information). He showed that each element could combine with multiple others to form different compounds, and always in whole number ratios.

This knowledge, taken together, provided the foundation for the chemical notation we use today. Chemical equations aren’t like mathematical equations, which have been around much longer. While the quantity of each element must be equal on both sides of the equation, you will never see an “equals” sign in a chemical equation. That’s because a chemical equation describes a process of change.

Comprehension Checkpoint

Chemical equations are written basically the same way today as they were in the 17th century.

Like all other reactions, the creation of the rusty muffler is an example of chemical change:

Fe + O2         Fe2O3
Reactants     Yield     Products

Again, iron and oxygen combine to form a specific iron oxide. The arrow indicates that this reaction proceeds to the right as it is written, meaning that iron oxide is formed. In some cases, the reaction can also go backward, and we use a double arrow showing that some reactions go in both directions. Unfortunately for the muffler, that’s not the case here.

Now look at the equation more closely: How many iron atoms are on the reactant side (left) versus the product side (right)? How many oxygen atoms? You will see that they’re not equal as the equation is currently written.

However, we know from the law of conservation of matter that atoms can’t be created or destroyed. In other words, we can’t just get rid of an iron atom or create an oxygen atom to make the equation work. Nor can we change the subscripts in the reaction, because doing so would suggest that we are starting with a different reactant or obtaining a different product.

What we can do is adjust the numbers of reactant and product components (iron atoms and/or oxygen molecules on the left side and iron oxide molecules on the right), because doing so doesn’t imply creating or destroying matter. It’s a way of bookkeeping – in order to produce a molecule that has two iron atoms, you need to find a source of those atoms.

Let’s see how this works with the reaction that creates rust.

We already know that the numbers of each type of atom aren’t equal on each side of the reaction. To address this, we can add coefficients in front of the reactants and products to adjust the number of particles and create a balanced equation. (If there is no coefficient, it means there is only one of that type of particle.)

Let’s look at the reaction atom by atom.

Fe + O2         Fe2O3
↑ 1 iron ↑ 2 iron

Since the rust molecule has two iron atoms, we must balance the Fe atoms by adding the coefficient “2” in front of the iron atom on the reactant side, since that is the only place where iron appears on the left side of the equation. Now we have two irons on each side of the equation.

2Fe + O2         Fe2O3
↑ 2 iron ↑ 2 iron

Moving on to the oxygen atoms, we find two on the left side of the equation, but three on the right side. We could mathematically balance the equation by using one and a half oxygen molecules, since each molecule is made up of two oxygen atoms. To better understand this, picture one oxygen molecule as two atoms bonded together O-O, therefore 1.5 oxygen molecules (O-O and O-O) provides three oxygen atoms:

2Fe + 1.5O2    Fe2 O3
↑ 3 oxygen ↑ 3 oxygen

In the real world, though, oxygen doesn’t occur in half-molecules. We can solve this conundrum by simply multiplying all the coefficients by two:

4Fe + 3O2    2Fe2O3
↑ 4 iron ↑ 6 oxygen ↑ 4 iron & 6 oxygen

Now we have 4 atoms of iron on the left side, and 4 (2 molecules, each containing 2 iron atoms) on the right side. For oxygen, there are 6 atoms on the left side (3 molecules of 2 atoms each) and 6 on the right side (2 molecules containing 3 oxygen atoms). Now we have a balanced equation.

Comprehension Checkpoint

To create a balanced equation, we can

Along with telling us exactly how much of a chemical compound is involved in a reaction, balanced chemical equations tell us is the proportions of “ingredients” required to make a particular product. It’s a bit like a recipe. Let’s say you’re making a batch of cookies. The recipe calls for:

  • 2 cups of flour
  • 1 cup of sugar

and promises you a batch of 12 cookies. You can follow the recipe and use 2 cups of flour and 1 cup of sugar and expect 12 cookies, or you can double the recipe and use 4 cups of flour and 2 cups of sugar and expect 24 cookies (or you can triple the recipe, or half it, or so on).

Similarly, to make a “batch” of 2 rust molecules, you need 4 iron atoms and 3 oxygen molecules:

4Fe + 3O2    2Fe2O3
↑ 4 iron ↑ 3 oxygen
molecules
↑ 2 rust
molecules

So, for every four atoms of iron and three molecules of oxygen, we get two molecules of rust. Much as you would do with a cookie recipe, you could double this: start with eight iron atoms and six oxygen molecules and make four molecules of rust. No matter how many times you multiply it, the base proportion always remains constant.

In the real world of chemistry, though, we don’t deal with individual atoms and molecules; there are a lot more than two molecules of rust on a muffler that falls off a car. That’s where the concept of the mole comes in handy. A mole is a quantity of particles, specifically a mole is 6.022 x 1023 particles. In other words, one mole of iron atoms contains 6.022 x 1023 atoms. Four moles of iron atoms contain four times that amount, or 28.088 x 1023 iron atoms. (To brush up on your mole math, see our module, The Mole and Atomic Mass).

Note that when we were balancing the equation, we were thinking about the numbers of individual atoms on each side, attending to the law of conservation of matter that we can’t create or destroy atoms. When we’re thinking about proportions, however, we’re thinking about the coefficients in front of the particles, which represent the number of moles of each type of particle consumed or produced. These coefficients tell us how many moles of product can be produced with the number of moles of reactant present at the beginning of the reaction.

A chemical equation is an ingeniously compact way of communicating a lot of information in a short sequence of parts. Modern chemical equations reflect our understanding of matter being composed of atoms and of chemical reactions as a process of breaking bonds and rearranging atoms into new compounds. Mass is conserved in a chemical reaction, and the number of particles on each side of the equation must reflect this. A balanced equation also communicates the proportions of products and reactants involved in that reaction.

Chemical equations are an efficient way to describe chemical reactions. This module explains the shorthand notation used to express how atoms are rearranged to make new compounds during a chemical reaction. It shows how balanced chemical equations convey proportions of each reactant and product involved. The module traces the development of chemical equations over the past four centuries as our understanding of chemical processes grew. A look at chemical equations reveals that nothing is lost and nothing is gained in a typical chemical reaction–matter simply changes form.

Robin Marks, M.A., Anthony Carpi, Ph.D. “Chemical Equations” Visionlearning Vol. CHE-4 (7), 2018.

Top


Page 6

Physical States and Properties

by Heather MacNeill Falconer, M.A./M.S., Gina Battaglia, Ph.D., Anthony Carpi, Ph.D.

If you’ve ever made cookies and left the kitchen door open, you’re probably aware that the aroma spreads throughout the house. It is strongest in the kitchen, where the cookies are baking, a little less in the dining or living room, and least in the upstairs corner bedroom. And if the door is closed in the corner bedroom, the cookie scent is even weaker.

This is a delicious example of diffusion, or the movement of matter from a region of high concentration (the cookie pan in the kitchen) to a region of low concentration (the corner bedroom). This principle of diffusion is fundamental throughout science, from gas exchange in the lungs to the spread of carbon dioxide in the atmosphere to the movement of water from one side of a cell’s plasma membrane to the other. However, the concept of diffusion is rarely as simple as molecules moving from one place to another. Temperature, the size of the molecules involved, the distance molecules need to travel, the barriers they may encounter along the way, and other factors all influence the rate at which diffusion takes place.

The universe is in constant motion: from the orbiting of planets around the sun, to the movement of particles from one area to another. And while on a grand scale it may appear that there is a rationale to this movement – for example, the planets in our solar system have regular revolutions that can be predicted – in truth there is a great deal of motion that occurs randomly.

When we learn about diffusion, we often hear about the movement of particles from an area of high concentration to an area of low concentration, as if the particles themselves are somehow motivated to move in this direction. But this movement is in fact a by-product of what scientists refer to as the “random walk” of particles. Molecules do not move in straight paths from Point A to Point B. Instead, they interact with their environment, bumping into other molecules and barriers encountered along their way, as well as interacting with the medium through which they are moving.

The observation of the spontaneous, random movement of small particles was first recorded in the first century BCE. Lucretius, a Roman poet and philosopher, described the dust seen in sunbeams coming through a window (Figure 1):

Two or more atoms that bond together form a(n)
Figure 1: Dust particles “dancing” in a ray of light. image © E.mil.mil

You will see a multitude of tiny particles mingling in a multitude of ways... their dancing is an actual indication of underlying movements of matter that are hidden from our sight... It originates with the atoms which move of themselves [i.e., spontaneously]… So the movement mounts up from the atoms and gradually emerges to the level of our senses, so that those bodies are in motion that we see in sunbeams, moved by blows that remain invisible.

While Lucretius’s “dancing” particles were likely dust particles or pollen grains that are affected by air currents and other phenomenon, his description is a wonderfully accurate assessment of what goes on at the molecular level. Many scientists have explored this random molecular motion in a variety of contexts, most famously by the Scottish botanist Robert Brown in the 19th century.

In 1828, while observing pollen granules suspended in water under a microscope, Brown discovered that the motion of the granules were “neither from currents in the fluid, nor from its gradual evaporation, but belonged to the particle itself.” After suspending various organic and inorganic substances in water and seeing this same inherent, random movement, he concluded that this random walk of particles – later termed Brownian motion in his honor – was a general property of matter that is suspended in a liquid medium. However, it would take nearly a century for scientists to mathematically quantify Brownian motion and demonstrate that this random movement of molecules dictates diffusion.

Comprehension Checkpoint

When molecules move from an area of high concentration to an area of low concentration,

About the same time that Brown was making his observations, a group of scientists including the French engineer Sadi Carnot and German physicist Rudolph Clausius were establishing a whole new field of scientific study: the field of Thermodynamics (see our Thermodynamics I module for more information). Clausius’s work in particular led to the development of the kinetic theory of heat – the idea that atoms and molecules are in motion and the speed of that motion is related to a number of things, including the heat of the substance. The molecules of a solid are generally considered to be locked in place (though they vibrate); however, the molecules of a liquid or a gas are free to move around, and they do: bumping in to one another or the walls of their container like balls on a pool table.

As molecules in a liquid or gas move through space, they bump into one another and follow random paths – moving in a straight line until something blocks their way and then bouncing off of that thing. This random molecular movement is constantly occurring and can be measured, giving a molecule’s mean free path – or, the average distance a particle moves between impacts with other particles.

It is this spontaneous and random motion that leads to diffusion. For example, as the scent molecules from baking cookies move into the air, they interact with air molecules – crashing into them and changing direction. Over time, these random processes will cause the scent molecules to disperse throughout the room. Diffusion is presented as a process in which a substance moves down a concentration gradient – from an area of high concentration to an area of low concentration. However, it is important to recognize that there is no directional force at play – the scent molecules are not pushed to the edge of the room because the concentration is lower there. It is the random movement of these molecules within the roomful of moving air molecules that causes them to evenly spread out throughout the entire space – bouncing off walls, moving through doors, and eventually moving through the whole house. In this way, it appears to move along a concentration gradient – from the kitchen oven to the most distant rooms of the house.

It may sound like a paradox – the movement of molecules are random, yet at the same time appear to occur along a gradient – but in practice, it’s actually quite logical. A simple illustration of this process can be seen using a glass of water and food coloring. When a drop of food coloring enters the water, the food coloring molecules are highly concentrated at the location where the dye molecules meet the water molecules, giving the water in that area a very dark color (Figure 2). The bottom of the glass initially has few or no food coloring molecules and so remains clear. As the food coloring molecules begin to interact with the water molecules, molecular collisions cause them to move randomly around the glass. As collisions continue, the molecules spread out, or diffuse, over space.

Two or more atoms that bond together form a(n)
Figure 2: Diffusion of a purple dye in a liquid.

Eventually, the molecules spread throughout the entire glass, becoming evenly distributed and filling the space. At this point, the molecules have reached a state of equilibrium in which no net diffusion is taking place and the concentration gradient no longer exists. In this state, the molecules are still moving haphazardly and colliding with each other; we just can’t see that motion because the water and color molecules are evenly dispersed throughout the space. Once equilibrium has been reached, the probability that a molecule will move from the top to the bottom is equal to the probability a molecule will move from the bottom to the top.

Comprehension Checkpoint

When equilibrium is achieved and molecules are equally distributed,

We know that diffusion involves the movement of particles from one place to another; thus, the speed at which those particles move affects diffusion. Since molecular motion can be measured by the heat of an object, it follows that the hotter a substance is the faster diffusion will take place in that substance. (Click the animation below to see how temperature affects diffusion.) If you were to repeat your food coloring and water experiment comparing a glass of cold to a glass of hot water, you would see that the color disperses much more quickly in the hot water. But what other factors influence the speed, or rate, at which diffusion takes place?

Two or more atoms that bond together form a(n)

Interactive Animation: The Effect of Temperature on Diffusion

In 1829, the Scottish physical chemist Thomas Graham first quantified diffusion behavior before the idea of atoms and molecules was widely established. Basing his observations on real-life “substances,” Graham measured the diffusion rates of gases through plaster plugs, fine tubes, and small orifices that were meant to slow down the diffusion process so that he could quantify it. One of his experiments, detailed in Figure 3, used an apparatus with the open end of a tube sitting in a beaker of water and the other end sealed with a plaster stopper containing holes large enough for gases to enter and leave the tube. Graham filled the open end of the tube with various gases (as indicated by the red tube in Figure 3), and observed the rate at which the gases effused, or escaped through the plaster plug. If the gas effused from the tube faster than the air outside of the tube moved in, the water level in the tube would rise. On the other hand, if the outside air moved through the plaster faster than the gas in the tube escaped to the outside, the water level in the tube would go down. He used the rate of change in the water level to determine the relative rate at which the different gases diffused into air.

Two or more atoms that bond together form a(n)
Figure 3: Thomas Graham's experiment to measure the diffusion rates of gases.

Graham experimented with many combinations of different gases and published his findings in an 1829 publication of the Quarterly Journal of Science, Literature, and Art titled “A Short Account of Experimental Researches on the Diffusion of Gases Through Each Other, and Their Separation by Mechanical Means.” He stated that when gases come into contact with each other, “indefinitely minute volumes” of the gases spontaneously intermix with each other until they reach equilibrium (Graham, 1829). However, he discovered that different types of gases did not mix at the same rate – rather, the rates at which two gases diffuse is inversely proportional to the square root of their densities, a relationship now known as Graham’s law. Although Graham’s original relationship used density, or mass per unit volume, the modern form of the equation uses molar mass, or the mass of one mole of a substance.

What Graham showed was that the molecular weight of a molecule directly affects the speed at which that molecule can move. Graham’s work actually helped lay the foundations of kinetic molecular theory because it recognized that at a given temperature, a heavy molecule would move more slowly than a light molecule. In other words, more kinetic energy is needed to move a large molecule at the same speed as a small molecule. You can think of it this way: A small push will get a tennis ball rolling quickly; however, it takes a much harder push to move a bowling ball at the same speed. At a given temperature, small molecules move faster, and will diffuse more quickly than large ones. View the animation below to see how atomic mass affects diffusion.

Two or more atoms that bond together form a(n)

Interactive Animation: The Effect of Atomic Mass on Diffusion

Comprehension Checkpoint

Rate of diffusion is influenced by

Graham later studied the diffusion of salts into liquids and discovered that the diffusion rate in liquids is several thousand times slower than in gases. This seems relatively obvious to us today, as we know that the molecules of a gas move faster and are more spread out than molecules in a liquid. Therefore, the movement of one substance within a gas occurs more freely than in a liquid. Diffusion in liquids is proportional to temperature, as it is in gases, as well as to the viscosity of the specific liquid into which the material is diffusing. (View the animation below to compare diffusion in gases and liquids.) Diffusion, in fact, can even take place in solids. While this is a very slow process, Sir William Chandler Roberts-Austen, a British metallurgist, fused gold plates to the end of cylindrical rods made of lead. He analyzed the lead rods after a period of 31 days and actually found that gold atoms had “flowed” into the solid rods.

While we have talked extensively about diffusion and concentration gradients, it was not until the mid-1800s when a German-born physicist and physiologist named Adolf Fick built upon Graham’s work and introduced the notion of a diffusion coefficient, or diffusivity, to characterize how fast molecules diffuse.

In his 1855 publication “On Diffusion” in Annalen der Physik, Fick described an experimental setup in which he connected cylindrical and conical tubes with solid salt crystals at the bottom to an “infinitely large” reservoir filled with freshwater (Figure 4). The solid salt crystals dissolved into the water in the tubes and diffused toward the water reservoir. A stream of freshwater swept the saltwater out of the reservoir. This stream of water kept the salt concentration at the very top of the tubes (the point where the salt solution met the water reservoir) close to zero. The dissolving salt at the bottom of the tube maintained a high salt concentration in the water at that end of the tube. Because the tubes had a different shape (conical versus cylindrical), the concentration gradient in the tubes differed, setting up a system in which diffusion could be compared in relation to a concentration gradient.

Two or more atoms that bond together form a(n)
Figure 4: Fick's experimental setup in which he connected cylindrical and conical tubes to a reservoir filled with freshwater. (Image from the 1903 publication, Collected Works, I. Stahel’sche Verlags-Anstalt, Würzburg: Germany.) image © Fick

Fick then calculated the diffusion rate of the salt by measuring the amount of salt that passed through the top of the respective tubes (just before they met the freshwater in the reservoir) within a given time period. He discovered that the movement rate of the salt solution into the water reservoir depended on the concentration difference between the solution at the bottom of the tube and the concentration of the solution leaving the tube and entering the reservoir. In other words – the higher the concentration of salt at the top of the tube, the faster it diffused into the water reservoir. You can see how concentration affects diffusion in the animation below.

Two or more atoms that bond together form a(n)

Interactive Animation: The Effect of Concentration on Diffusion

After studying the phenomenon, Fick hypothesized that the relationship between the concentration gradient and the diffusion rate was similar to what Joseph Fourier, a French mathematician and physicist, found in his study of heat conduction in 1822. Fourier had described the rate of heat transfer through a substance as proportional to the difference in temperature between two regions. Heat moves from warmer to cooler objects, and the greater the temperature difference between the two objects, the faster the heat moves. (This is why your mug of hot coffee cools off much faster outside on a cold morning than when you leave it in your heated apartment). Using Fourier’s law of thermal conduction as a model, Fick created a mathematical framework for the movement of salt into the water, proposing that the diffusion rate of a substance is proportional to the difference in concentration between the two regions. What this means for diffusion of a substance is that if the concentration of a given substance is high in relation to the substance it is diffusing into (e.g., food coloring into water), it will diffuse faster than if the concentration difference is low (e.g., food coloring into food coloring). The application of a successful principle from one branch of science to another is not uncommon, and Fick was a classic example of this process. Fick knew of Fourier’s work because he had modeled his experimental apparatus on that of Fourier. Thus it was natural for him to apply Fourier’s law to diffusion. While he had no way to know that the underlying mechanism of heat conduction and diffusion were both based on atomic collisions (in fact, some researchers at the time still doubted the existence of atoms), he had a feeling. That feeling, and the existence of atoms themselves, would be mathematically proven some 50 years later when Albert Einstein published his seminal work, Investigations on the Theory of the Brownian Movement (Einstein, 1905).

The diffusion coefficient, or diffusivity D, defined by Fick is a proportionality constant between the diffusion rate and the concentration gradient. The diffusion coefficient is defined for a specific solute-solvent pair, and the higher the value for the coefficient, the faster two substances will diffuse into one another. For example, at 25°C the diffusivity of gaseous air into gaseous water is 0.282 cm2/sec (Cussler, 1997). At the same temperature, the diffusivity of dissolved air into liquid water is 2.00 x 10-5 cm2/sec, a much lower number than that for the two gases, representing the much slower diffusion rate in liquids compared to gases. And the diffusivity of dissolved helium into liquid water at 25°C is 6.28 x 10-5 cm2/sec – higher than that of dissolved air, representing the smaller size of helium atoms compared to the nitrogen and oxygen molecules in air.

Comprehension Checkpoint

When dissolved into water, helium has a higher diffusivity than air. This means that helium dissolves at a ______ rate than air.

Yet another factor that influences the rate at which diffusion occurs is the distance a molecule travels before bumping into something (referred to as a molecule’s mean free path). Imagine taking a container filled with a gas and putting it under pressure so that the molecules in the gas are squeezed together. This would slow the rate of diffusion through the gas because the molecules travel a shorter distance before colliding with something else and changing direction. (The animation below shows the effect of pressure on diffusion.)

Two or more atoms that bond together form a(n)

Interactive Animation: The Effect of Pressure on Diffusion

This is an important factor affecting the difference in diffusion rates in gases versus liquids versus solids; because gas particles are the most spread out of the three, molecules travel the furthest between collisions and diffusion occurs most rapidly in this state (Figure 5).

Two or more atoms that bond together form a(n)
Figure 5: The three states of matter at the atomic level: gas, liquid, and solid.

To fully understand why we can smell the cookies baking in the kitchen from the bedroom we also have to consider another process at work here – advection. Advection involves the transfer of a material or heat due to the movement of a fluid. So, because people walk through the rooms of your house and because heat rises from your radiators, the air is constantly moving, and that movement carries and mixes the scent molecules in your house. In many situations (such as your house), the effects of advection exceed those of diffusion, but these processes work in tandem to bring you the cookie smell.

From the traveling smells of cookies to the dissolving of salt into water, diffusion is a process happening around (and within!) us every second of every day. It is a process that is critical to moving oxygen across the membranes of our lungs, moving nutrients through soil to be taken up by plants, dispersing pollutants that are released into the atmosphere, and a whole host of other events that are necessary for life to exist.

The process of diffusion is critical to life, as it is necessary when our lungs exchange gas during breathing and when our cells take in nutrients. This module explains diffusion and describes factors that influence the process. The module looks at historical developments in our understanding of diffusion, from observations of “dancing” particles in the first century BCE to the discovery of Brownian motion to more recent experiments. Topics include concentration gradients, the diffusion coefficient, and advection.

Key Concepts

  • Diffusion is the process by which molecules move through a substance, seemingly down a concentration gradient, because of the random molecular motion and collision between particles.

  • Many factors influence the rate at which diffusion takes place, including the medium through with a substance is diffusing, the size of molecules diffusing, the temperature of the materials, and the distance molecules travel between collisions.

  • The diffusion coefficient, or diffusivity, provides a relative measure at specific conditions of the speed at which two substances will diffuse into one another.

  • HS-C5.4, HS-PS3.A3, HS-PS3.B5
  • Brown, R. (1828). A brief account of microscopical observations made on the particles contained in the pollen of plants. Philosophical Magazine, 4, 161-173.
  • Cussler, E.L. (1997). Diffusion: Mass transfer in fluid systems (2nd ed). New York: Cambridge University Press.
  • Einstein, A. (1956). Investigations on the theory of Brownian movement. New York: Dover.
  • Fick, A. (1855). Ueber Diffusion [On diffusion]. Annalen der Physik und Chemie von J. C. Pogendorff, 94, 59-86.
  • Graham, T. (1829). A short account of experimental researches on the diffusion of gases through each other, and their separation by mechanical means. Quarterly Journal of Science, Literature, and Art, 2, 74–83.
  • Lucretius. (1880). On the nature of things. (trans. J.S. Watson). London: George Bell & Sons.

Heather MacNeill Falconer, M.A./M.S., Gina Battaglia, Ph.D., Anthony Carpi, Ph.D. “Diffusion I” Visionlearning Vol. CHE-3 (4), 2015.

Top


Page 7

Physical States and Properties

by Robin Marks, M.A., Anthony Carpi, Ph.D.

This is an updated version of the Water module. For the previous version, go here.

Before we start, get yourself a glass of water. By the time you’ve reached the end, you’ll have a much greater appreciation for this miracle liquid.

Got your glass? Now take a sip and think about all the roles water plays in your life. For one thing, your body can’t function more than a few days without it. You use water to wash yourself, your clothes, and your car. Water puts out fires, cooks our food, makes our soap get sudsy, and hundreds of other things. Water is absolutely essential to our lives on Earth.

Water is so central to our existence that you might be surprised to learn that it’s a rare and unusual substance in the universe. Water is at once so vital and so scarce that exobiologists (scientists looking for life beyond Earth) set their sights on planets where water might exist. Life, it seems, can tough it out in acid, lye, extreme salt, extreme heat, and other conditions that would kill us humans. But it can’t exist without water.

Despite its scarcity across the universe, water is so abundant on Earth that we aren’t always aware of how special it is. For starters, water is the only substance that exists naturally on our planet as a solid (ice and snow), liquid (rivers, lakes, and oceans), and a gas (water in the atmosphere as humidity). As you might recall (or can read about in our module on States of Matter), water molecules are in a different energy state in each phase. The amount of energy required to go from solid to liquid and liquid to gas is related to how water molecules interact with each other. Those interactions are, in turn, related to how the atoms within a water molecule interact with each other.

Our Chemical Bonding: The Nature of the Chemical Bond module discussed how a dipole forms across a water molecule; in the bond between oxygen and hydrogen, the electrons are shared unequally, drawn a bit more to the oxygen. As a result, a partial negative charge (ð-) forms at the oxygen end of the molecule, and a partial positive charge (ð+) forms at each of the hydrogen atom ends (Figure 1).

Two or more atoms that bond together form a(n)
Figure 1: The dipoles arise in a water molecule because of unequal sharing of electrons.

Since the hydrogen and oxygen atoms in the molecule carry opposite (though partial) charges, nearby water molecules are attracted to each other like tiny little magnets. The electrostatic attraction between the ð+ hydrogen (ð stands for partial charge, a value less than the charge of an electron) and the ð- oxygen in adjacent molecules is called hydrogen bonding (Figure 2).

Two or more atoms that bond together form a(n)
Figure 2: Hydrogen bonds between water molecules. The slight negative charge on the oxygen atom is attracted to the slight positive charge on a hydrogen atom.

Hydrogen bonds make water molecules "stick" together. These bonds are relatively weak compared to other types of covalent or ionic bonds. In fact, they are often referred to as an attractive force as opposed to a true bond. Yet, they have a big effect on how water behaves. There are many other compounds that form hydrogen bonds, but the ones between water molecules are particularly strong. Figure 2 shows why. If you look at the central molecule in this figure you see that the oxygen end of the molecule forms hydrogen bonds with two other water molecules; in addition, each hydrogen on the central molecule is attracted to a separate water molecule. As the illustration shows, each water molecule forms attractions with four other water molecules, a network of connections that makes the hydrogen bonding in water particularly strong and lends the substance its many unique properties.

Now it’s time to make use of that glass of water. If you have some ice cubes, drop one in your glass. You’ll notice that it floats. Its ability to bob to the top of the water line means that the ice (water in its solid state) is less dense than liquid water. (To review density and buoyancy, see our Density module) This isn’t a common state of affairs; if you put a chunk of solid wax into a vat of molten wax, it will sink toward the bottom (and possibly melt before it gets there).

To understand what causes ice to float but solid wax to sink, let’s think first about what happens when a liquid turns to a solid (again, the States of Matter module can be a handy review here). In a liquid, the molecules have enough kinetic energy to keep moving around. As molecules come near to each other, they are drawn together by intermolecular forces. At the same time, molecules have enough kinetic energy to break free of those forces and be drawn to other nearby molecules. Thus the liquid flows because intermolecular attractions can be broken and reformed.

A liquid freezes when the kinetic energy is reduced (i.e. the temperature is reduced) enough that the attractive forces between molecules can no longer be broken, and the molecules become locked in a static lattice. For nearly all compounds, the lower energy and lack of movement between molecules means the molecules in a solid are packed together more tightly than the liquid state. This is the case with wax and so solid wax is denser than the liquid and sinks.

In the case of water, though, the shape of the molecule and the strength of the hydrogen bonds affect the arrangement of the molecules. In liquid water, hydrogen bonding pulls molecules closely together. As water freezes, the dipole ends with like charges repel each other, forcing the molecules into a fixed lattice in which they are farther from each other than they are in liquid water (Figure 3). More space between molecules makes the ice less dense than liquid water, and thus it floats.

Two or more atoms that bond together form a(n)
Figure 3: When water freezes, the similarly-charged ends of the dipoles repel each other, pushing molecules apart. This means there is more space between molecules in the solid than in the liquid, making the solid (aka, ice) less dense.

Water is sometimes referred to as the “universal solvent,” because it dissolves more compounds than any other liquid known. The polarity of the water molecule allows it to readily dissolve other polar molecules, as well as ions. (See our Solutions, solubility, and colligative properties module for a deeper discussion of dissolution.)

This ability to dissolve substances is one of the properties that makes water vital for life. Most biological molecules, such as DNA, proteins, and vitamins are polar, and important ions such as sodium and potassium are also charged. In order for any of these compounds to carry out functions in the body, they have to be able to circulate in the blood and the fluid within and between cells, all of which are mostly water. Because of its polarity, water is able to dissolve these and other substances, allowing their free movement around the body. A few biomolecules, such as fats and cholesterol, aren’t polar, and don’t dissolve in water – however, the body has developed unique ways to circulate and store these substances.

Water is also able to dissolve gasses such as oxygen, allowing fish, plants, and other aquatic life to access this dissolved oxygen (Figure 4). O2 isn’t a polar molecule; it dissolves because the polar charges in the water molecule induce a dipole in the oxygen, making it soluble and so available to aquatic life. (Learn more about induced-dipole interactions in our Properties of Liquids module.)

Two or more atoms that bond together form a(n)
Figure 4: When water and oxygen molecules meet (left), the negative dipole of water repels electrons around the oxygen molecule, creating a temporary dipole in the oxygen molecule (right).

Let’s return to your water glass. Fill the glass just to the rim and stop. Then, slowly, add a little bit more. You’ll see that you can actually fill the glass a bit past its rim, and the edges of the water will round out against the glass, holding the water in.

Once again, hydrogen bonding is behind this act, resulting in cohesion. Cohesion occurs when molecules of the same kind are attracted to each other. In the case of water, the molecules form strong hydrogen bonds, which hold the substance together. As a result, water is highly cohesive, in fact, it is the most cohesive of all non-metallic liquids.

Cohesion occurs throughout your glass of water, but it’s especially strong at the surface. Molecules there have fewer neighbors (because they have none at the very surface), and so create stronger bonds with the molecules that are near them. The result is called surface tension, or the ability of a substance to resist disruption to its surface. Dip your finger into your water glass and then pull it out. The drop that forms at the end of your fingertip is held together by surface tension.

Surface tension was the misunderstood central player in a raucous debate between Galileo Galilei and his chief rival, Ludovico delle Colombe in 1611. Delle Colombe, a philosopher, was at odds with some of Galileo’s ideas, including his explanation that ice floats on water because it is less dense. So the philosopher challenged Galileo to a debate, which delle Colombe believed would prove his own intellectual superiority.

Delle Colombe championed the (incorrect) idea that ice floats not because of density, but because of its shape, which he saw as broad and flat, as is ice on a lake. To prove the “truth” of his theory, he used ebony wood, which is slightly denser than water, in a demonstration before an audience of curious spectators. He dropped a sphere of the wood into water, and it sank. He then placed a thin wafer of the wood flat on the water’s surface, and it floated. Delle Colombe pronounced himself the winner.

Galileo left frustrated. His observations of the world gave him evidence that his explanation, not delle Colombe’s, was right, but he couldn’t explain the outcome of delle Colombe’s experiment.

Had he known about molecules and dipoles and hydrogen bonds at the time, Galileo certainly would have offered this explanation: When delle Colombe floated the thin ebony disc, he was taking advantage of the cohesive nature of water and the surface tension that arises from it (Figure 5). As the ebony wafer appeared to float on the water, the force exerted by its mass was distributed throughout the surface of the water beneath it. In other words, a single pinpoint-sized area of surface water only had to support the pinpoint-sized piece of ebony just above it. The hydrogen bonds between the water molecules were strong enough to support the weight of the disc. When delle Colombe placed the sphere in the water, however, the pinpoint-sized area that first touched the water bore the weight of the entire sphere, which was more than the water’s surface tension could support. Had Galileo known this at the time, he could have disproved delle Colombe easily – had he simply pushed the wafer through the surface to break the surface tension, the wafer would have sunk.

Two or more atoms that bond together form a(n)
Figure 5: Water molecules at the surface form stronger hydrogen bonds between them than do molecules in the rest of the water. These stronger bonds are responsible for surface tension. image © USGS

This same surface tension is what allows leaves to stay at the surface of a lake and dewdrops to adhere to a spider’s web. Even some animals take advantage of this phenomenon – the Basilisk lizard (Figure 6), water striders, and a few other small animals and bugs appear to "walk" on water by taking advantage of the surface tension of water.


Figure 6: A Basilisk lizard (Basiliscus basiliscus) runs on the water surface. Movie S1 from Minetti A, Ivanenko Y, Cappellini G, Dominici N, Lacquaniti F. "Humans Running in Place on Water at Simulated Reduced Gravity". PLOS ONE. DOI:10.1371/journal.pone.0037300

For your next observation, take another sip of water, and notice the side of the glass. Chances are you’ll see a few drops stuck to it. Gravity is pulling down on these drops, so something else must be keeping them stuck there. That something else is adhesion, the attraction of water to other kinds of molecules; in this case, the molecules that make up the glass. Because of the polarity of the molecule, water exhibits stronger adhesion to those surfaces that have some net electrical charge, and glass is one such surface. But place a drop of water on a non-polar surface, such as a piece of wax paper and you will see it take a different shape than one to which it adheres. On the wax paper, the water droplets take the shape of a true droplet because there is little adhesion and the cohesive forces pull the drop into a sphere. But on glass you will see the droplets flatten and deform a bit as the adhesive forces draw it more to the surface of the glass.

Both cohesion and adhesion (Figure 7) occur with many compounds besides water. Pressure sensitive tapes, for example, stick to surfaces because they are coated with a high viscosity fluid that adheres to the surface to which they are pressed. Generally, you can overcome this adhesive force by pulling, for example – you can easily lift a Post-it® Note from a page. But sometimes the adhesive forces are stronger than the forces holding the surface together – pull tape off of a piece of paper and you remove pieces of the paper with the tape.

Let’s return to our glass of water, and look inside to where the water surface meets the glass. The very edge of the water surface curves upward slightly on the glass. That’s also adhesion – the water is drawn up the surface by adhesion with the glass. If you have a clear plastic straw, you can put one end of it into the water and see that the liquid climbs up the straw a bit, above the surface of the remaining glass of water. It’s actually moving upward against gravity!

What’s happening in your straw is a phenomenon called capillary action (Figure 7). Capillary action occurs in small tubes, where the surface area of the water is small, and the force of adhesion—water’s attraction to the polar glass or other material—overcomes the force of cohesion between those surface molecules.

Two or more atoms that bond together form a(n)
Figure 7: The attraction of water molecules to the sides of a narrow vessel (adhesion, red arrows) is stronger than the cohesion (orange arrows) drawing water molecules together. The result is capillary action, in which the force of adhesion pulls the fluid upwards (purple arrows).

Another way to see the effects of adhesion and cohesion is to compare the behavior of polar and nonpolar liquids. When you put water in a test tube, adhesion makes the water along the edges move slightly upward and creates a concave meniscus. Liquid mercury, on the other hand, is not polar and therefore not attracted to glass. In a test tube, cohesion at the surface of the mercury is much stronger than adhesion to the glass. The surface tension in the mercury forms a convex meniscus, much the same as the way water forms a slight bulge over the top of your very full glass (Figure 8).

Two or more atoms that bond together form a(n)
Figure 8: Water and mercury behave differently in a test tube made of polar glass. Water adheres to the glass, bringing the sides upwards and forming a concave surface. Nonpolar mercury is not attracted to the glass. Cohesion between the mercury atoms creates surface tension that forms a convex surface. image © USGS

Adhesion and capillary action are among the forces at play that help plants take up water (and dissolved nutrients) in their roots. Capillary action also keeps your eyes from drying out, as saline water flows from tiny ducts in the outer corners of your eyes. With each blink, you spread the water away from the duct, and capillary action brings more fluid to the surface.

If you want to see capillary action at work, put a few drops of red food coloring in your glass of water, and then drop a stalk or two of leafy celery into it. After a day or two, your green celery will be streaked with red.

Water is a truly unusual and important substance. The unique chemical properties of water that give rise to surface tension, capillary action, and the low density of ice play vital roles in life as we know it. Floating ice protects aquatic organisms and keeps them from being frozen in the winter. Capillary action keeps plants alive. Surface tension allows lily pads to stay on the surface of a lake. In fact, water’s chemistry is so complex and important that scientists today are still striving to understand all the feats this simple substance can perform.

Key Concepts

  • Water has a number of unique properties that make it important in both the chemical and biological worlds.

  • The polarity of water molecules allows liquid water to act as a "universal solvent," able to dissolve many ionic and polar covalent compounds.

  • The polarity of water molecules also results in strong hydrogen bonds that give rise to phenomena such as surface tension, adhesion, and cohesion.

  • Everts, S. (2013). Galileo on ice: Researchers commemorate the scientist's debate on why ice floats on water. Chemical and Engineering News, 91(34), p.28-29.

  • Heilbron, J.L. (2012). Galileo. Oxford University Press.
  • Lo Nostro, P. and Ninham, B.W. (2014). Aqua incognita: Why ice floats on water and Galileo 400 years on. Connor Court Press.
  • MachLachlan, J. (1999). Galileo Galilei: First physicist. Oxford University Press.
  • Whitehouse, D. (2009). Renaissance genius: Galileo Galilei and his legacy to modern science. Sterling Press.

Robin Marks, M.A., Anthony Carpi, Ph.D. “Water” Visionlearning Vol. CHE-4 (6), 2018.

Top


Page 8

It’s nearly the start of the school year, and you’ve gathered 10 friends for an end-of-semester bonfire. What would a bonfire be without s’mores? You pack supplies, making sure you have enough to make one s’more for everyone.

Two or more atoms that bond together form a(n)

You can think of making a s’more as a chemical equation (see our Chemical Equations module for more on those):

2 Graham crackers + 1 piece of chocolate + 4 mini-marshmallows → 1 s’more

Just as with a chemical equation, the coefficients in front of the “reactants” and “products” show the proportions in which they react to produce the desired product—one s’more.

So, to make 10 s’mores, you would need:

20 Graham crackers + 10 pieces of chocolate + 40 mini-marshmallow → 10 s’mores

Congratulations, you’ve just made it through your first exercise in what chemists call stoichiometry. This mouthful of a term was coined in the 1790s by chemist Jeremias Benjamin Richter, who became fascinated by the proportional mathematics of combining chemicals, convinced that it held clues to the nature of matter (which it does indeed; Dalton drew on this math to devise his early atomic theory, described in depth in our module Early Ideas about Matter: From Democritus to Dalton. Richter combined the Greek words stoicheion, which means “element,” and metron, which means “measure.” In other words, stoichiometry is a way of measuring the amount of each reactant combining in a chemical reaction, in our case, the amounts of reactant (20 graham crackers, 10 pieces of chocolate and 30 mini-marshmallows) that in turn predict the amount of product (10 s’mores), or vice versa.

Stoichiometry may seem like a complicated word, but it’s a fairly straightforward concept when you apply it to chemical equations: the proportions expressed in a chemical equation (the coefficients) can be used to predict how much product will be produced from a given measure of reactants.

For example, we used stoichiometry to determine how many s’more “reactants” we would need to make 10 s’mores. We can also use stoichiometry to predict how much product we’ll get with the amount of each reactant we have. If we have lots and lots of chocolate and marshmallows but only 12 graham crackers, how many s’mores can we make?

Again, our equation is: 2 Graham crackers + 1 piece of chocolate + 4 mini-marshmallows → 1 s’more

If we have 12 graham crackers, that’s enough to make 6 s’mores. It doesn’t matter how much extra chocolate we have, because without the graham crackers, it isn’t a s’more.

So the mole ratio of graham crackers to s’mores produced is:

Using the same concept of mole ratios as explained above, stoichiometry is used to figure out how much reactant is needed to make a desired quantity of product in a laboratory or manufacturing facility. An important industrial example is the production of nitrogen-based fertilizer, which provides important nutrients to the soil and allows modern farmers to grow more food per acre.

For centuries, farmers have understood the importance of adding nutrients to the soil in which they grow crops, but prior to the 1900s they were limited to using animal manure or expensive, naturally occurring mineral deposits as fertilizer. In the 1840s, the German chemist Justus von Liebig identified nitrogen as fertilizer’s key ingredient. However, despite the abundance of nitrogen in the atmosphere, there was no easy way to convert nitrogen to a form that could be taken up by plants.

This all changed in the early 1900s when the German chemist Fritz Haber invented a chemical process for converting nitrogen to ammonia (NH3), the compound that often gives household cleaners their characteristic smell, and which plants can use as a source of nitrogen. His initial method was only economical on a small scale, so Haber worked with a German colleague, Carl Bosch, to adapt this process to work at an industrial scale. The Haber-Bosch process is sometimes referred to as one of the most significant inventions of the 20th century, and it led to Haber winning the Nobel Prize in Chemistry in 1918. In its equation form, the Haber-Bosch process is relatively simple:

N2 + 3H2 → 2NH3

The ability to perform this simple reaction on a large scale had important historical consequences. Cheap ammonia provided an avenue to widely available inexpensive fertilizers, which created a boom in agriculture (and an associated increase in population) in the 20th century. And it indirectly prolonged World War I by providing Germany with an inexpensive source of the nitrogen necessary to make gunpowder. Some scientists have more recently questioned whether the Haber-Bosch process is a sustainable practice, given the environmental impact of agriculture and a growing population, as well as the fact that considerable energy is required to generate the hydrogen gas.

Two or more atoms that bond together form a(n)
Spraying rice fields with fertilizer. The manufacturing of nitrogen-based fertilizers relies on stoichiometry to calculate how much starting material (N2 and H2) is needed to produce the desired amount of ammonia (NH3) to be used in the fertilizer. image © Jan Amiss

Let’s apply our stoichiometry discussion here and imagine that an agricultural company needs to manufacture 1,500 kilograms of NH3 to meet the demand for fertilizer. How much N2 and H2 would they need to start with?

Again, the equation is: N2 + 3H2 → 2NH3

Looking at the equation, we see that the mole ratio of N2 required to produce NH3 is:

(1 mol N2) / (2 mol NH3)

Now we’ll use this mole ratio to determine how much reactant we need to start with to make 1,500 kilograms of ammonia.

First, a reminder: whenever we are calculating amounts of substance in a reaction, we have to convert the mass of each substance into moles. Why? Because the substances involved don’t have equal weights. Think of it in terms of the s’more: 1 piece of chocolate weighs a lot more than 1 mini-marshmallow. If we used mass in the equation instead of number of pieces, we might say that one s’more requires 1 gram of chocolate and 4 grams of mini-marshmallows. But in reality, that would amount to one piece of chocolate and about 50 mini-marshmallows! (For more on converting from grams to moles, see our module about the mole and atomic mass.)

So, we know that we want to make 1500 kg of NH3, Let’s start by converting kilograms to grams as follows

1,500 kg of NH3 x 1000 g/kg = 1,500,000 g of NH3

Then, we need to calculate how many moles that is. To do that, we multiply the molecular mass of NH3 (17 g per mole) by the number of grams, setting the equation up so that the grams cancel and the answer is in moles. We see that:

1,500,000 g NH3 × (1 mol NH3) / (17 g NH3) = 88,325 mol NH3

Next, we can use the mole ratio to figure out how many moles of N2 will be needed. Since we need 1 mole of N2 to produce 2 moles of NH3, we use that mole ratio to determine how many moles of N2 will be needed to produce 88,235 moles of NH3:

88,235 mol NH3 × (1 mol N2) / (2 mol NH3) = 44,117 mol N2

Now, use the molecular mass of N2 to figure out how many grams of N2 are required, then convert the grams of N2 to kg of N2, since that’s the units we want for our answer:

44,117 mol N2 × (28 g N2) / (1 mol N2) = 882,340 g N2, or 882.340 kg

Now that we know how many kilograms of N2 we would need, we can use the mole ratio of the reactants (N2 and H2) to figure out how many moles of H2 are required.

Remember, the equation states:

N2 + 3H2 → 2NH3

The mole ratio for H2 to N2 is 3 to 1. So, for every mole of nitrogen, we’ll need three times as many moles of hydrogen:

(3 mol H2) / (1 mol N2 )

Remember, we need to calculate how many moles of hydrogen are needed, and then convert those moles of hydrogen into grams of hydrogen. Using the moles of nitrogen we calculated above, we can get kg of H2:

44,117 mol N2 x (3 mol H2) / (1 mol N2 ) × (2 g H2) / (1 mol H2) = 264,702 g H2, or 264.702 kg

This is important information for the fertilizer manufacturer. While nitrogen is readily available from the air, hydrogen gas is not. So the manufacturer would likely have to purchase the hydrogen gas, which is expensive to generate, potentially explosive, and difficult to transport and store. Therefore, the manufacturer needs to know precisely how much hydrogen gas is required.

In the case above, the manufacturer will have an unlimited amount of nitrogen gas, but a precise amount of hydrogen gas. Therefore, the amount of hydrogen gas will limit the amount of ammonia that can be made (just like the number of graham crackers can limit the number of s’mores that can be made).

We would say that hydrogen is the limiting reactant, meaning that this is the reactant that will be used up first. As a result, the amount of it will determine how much product is produced. Determining how much reactant is required to produce a specific amount of product is one of the most important applications of stoichiometry.

We’ll illustrate this first with the s’mores. Let’s say you have the following amounts of s’more “reactants”:

120 graham crackers

70 pieces of chocolate

200 mini-marshmallows

Here, again, is the s’mores equation: 2 Graham crackers + 1 piece of chocolate + 4 mini-marshmallows → 1 s’more

How many s’mores can you make from your reactants? That will depend on the limiting reactant, the one which will run out first. To determine with reactant is limiting, you will first need to calculate how many s’mores you can make with each of the reactants. You can do this using the mole ratio:

Limiting reactant is an important concept in any manufacturing process. A manufacturer knows they want to make a certain amount of a specific product, and will purchase the reactants accordingly. In many cases, it is more economical to make the most expensive reactant be the limiting one, reducing the cost of excess and waste.

Silver nitrate is a good example. This compound, AgNO3, has been used since ancient times as a disinfectant and wound-healing agent. Today it is used in bandages and other medical applications, as well as water purification. It can be easily made by reacting pure silver with nitric acid, according to the equation:

3Ag + 4HNO3 → 3AgNO3 + 2H2O + NO

Silver is a much more expensive reactant than nitric acid, so someone using it to produce silver nitrate will probably want to make silver the limiting reactant.

By starting with a set amount of each reactant, you can determine not only the limiting reactant but also the mass of product that will be produced and the amount of reactant that remains in excess.

Let’s say we start with 150g of silver and 150 g of nitric acid. How much AgNO3 can we make, and which reactant is the limiting one?

To find the answer takes a few steps:

1) convert each reactant to moles 2) use the mole ratio to determine how many moles of one reactant would be required to use up the other 3) calculate the amount of product based on using up all the limiting reactant.

Step 1: Converting to moles

150g Ag × (1 mole) / (108 g Ag) =1.39 mol Ag

and

150g HNO3 × (1 mole) / (63 g HNO3) = 2.38 mol HNO3

Step 2: Using the mole ratio to equate moles of Ag to moles of HNO3

(4 mol HNO3) / (3 mol Ag)

Since silver is our expensive reactant, we want to use it all up. We can calculate how many moles of HNO3 is required to react with the whole 1.39 mol of Ag, setting up the equation so that moles of HNO3 cancel:

1.39 mol Ag × (4 mol HNO3) / (3 mol Ag) = 1.85 mol HNO3 required

To use up all the, Ag, we need 1.85 moles of HNO3. Look at our calculations above. How much HNO3 do we have? We have 2.38 moles – more than we need. In other words, if we put all of both reactants together, the silver will be used up first, and there will be HNO3 left over. That makes silver the limiting reactant.

This example shows the importance of converting to moles first. We started with the same mass of each reactant, 150 g. But mass doesn’t tell us how many particles there are. That is what the unit of moles tells us.

Knowing that silver is the limiting reactant, we can go further and determine how many moles of AgNO3 is produced from the 1.39 moles of Ag we are starting with. This time we use the mole ratio between Ag and AgNO3.

In the reaction:

3Ag + 4HNO 3 → 3AgNO3 + 2H2O + NO

There are 3 moles of AgNO3 produced for every 3 moles of Ag used. So:

1.39 mol Ag × (3 mol AgNO3) / (3 mol Ag) = 1.379 mol AgNO3 produced

Calculations such as these are vital to our ability to manufacture and use chemicals efficiently, as well as to our ability to understand the impact of the reactions that take place in our everyday world. For example, an engineer for a paint manufacturer must consider the mole ratios of different chemicals in the paint, which will determine the cost of producing that paint. On a grander scale, stoichiometry plays a role in understanding climate change: if we know the quantities of different types of fossil fuels burned in a year, we can determine how much CO2 has been added to the atmosphere. From planning for s’mores to streamlining manufacturing and generating environmental data, we can use stoichiometry to predict and plan the outcome of many chemical processes.

Stoichiometry is the mathematics of chemistry. Starting with a balanced chemical equation, we make use of the proportional nature of chemical reactions to calculate the amount of reactant needed at the start or predict the amount of product that will be produced. While it may not seem all that “chemical,” stoichiometry is a concept that underlies our ability to understand the impact and implications of many chemical processes. A bandage manufacturer may use mole ratios to determine how much silver is required (and therefor the cost) to treat a batch of bandages with silver nitrate. A fertilizer company might apply the concept of limiting reactant to figure out how much product they can produce with a given amount of hydrogen gas. And so on. Stoichiometry, mole ratios, and limiting reactants are indispensable concepts for fully understanding any chemical process.

  • Stoichiometry uses the proportional nature of chemical equations to determine the amount of reactant needed to produce a given amount of product or predict the amount that will be produced from a given amount of reactant.

  • The mole ratio shows the proportion of one reactant or product in a reaction to another, and is derived from the balanced chemical equation. While we may need to adjust the amount of reactants to yield more product, the ratio of reactants to products is always the same as the balanced reaction.

  • The limiting reactant is the chemical used up first in a reaction. it can be determined by comparing the number of moles of each reactant on hand and the mole ratio between reactants and products in the balanced reaction.

Stoichiometry is the mathematics of chemistry. Starting with a balanced chemical equation, we make use of the proportional nature of chemical reactions to calculate the amount of reactant needed at the start or predict the amount of product that will be produced. While it may not seem all that “chemical,” stoichiometry is a concept that underlies our ability to understand the impact and implications of many chemical processes. A bandage manufacturer may use mole ratios to determine how much silver is required (and therefor the cost) to treat a batch of bandages with silver nitrate. A fertilizer company might apply the concept of limiting reactant to figure out how much product they can produce with a given amount of hydrogen gas. And so on. Stoichiometry, mole ratios, and limiting reactants are indispensable concepts for fully understanding any chemical process.

  • The Haber-Bosch Reaction: An Early Chemical Impact On Sustainability
  • https://pubs.acs.org/cen/coverstory/86/8633cover3box2.html
  • Overview of the Haber-Bosch Process:

  • https://www.thoughtco.com/overview-of-the-haber-bosch-process-1434563

Robin Marks, M.A., Anthony Carpi, Ph.D. “Stoichiometry” Visionlearning Vol. CHE-4 (8), 2019.


Page 9

Atomic Theory and Structure

by Adrian Dingle, B.Sc., Anthony Carpi, Ph.D.

This is an updated version of our Atomic Theory I module. For the previous version, please go here.

By the late 1800’s, John Dalton’s view of atoms as the smallest particles that made up all matter had held sway for about 100 years, but that idea was about to be challenged. Several scientists working on atomic models found that atoms were not the smallest possible particles that made up matter, and that different parts of the atom had very distinct characteristics.

The English scientist Michael Faraday can reasonably be considered one of the greatest minds ever in the fields of electrochemistry and electromagnetism. Somewhat paradoxically, all of Faraday’s pioneering work was carried out prior to the discovery of the fundamental particle that these electrical phenomena depend upon. However, one of Faraday’s earliest experimental observations was a crucial precursor to the discovery of the first subatomic particle, the electron.

As early as the mid-17th century, scientists had been experimenting with glass tubes filled with what was known then as rarefied air. Rarefied air referred to a system in which most of the gaseous atoms had been removed, but where the vacuum was not complete. In 1838, Faraday noted that when passing a current through such a tube, an arc of electricity was observed. The arc started at the negative plate (known as the cathode) and traveled through the tube to the oppositely charged anode (Faraday, 1838).

In his experiments, Faraday observed a luminescence that started part way down the tube, and traveled toward the anode. This left an area between the cathode and the start of the luminescence that was not illuminated, and subsequently became known as Faraday’s dark space (Figure 1). Faraday couldn’t fully explain his observations, and it took a number of further developments in terms of the technology of the tubes, before a greater understanding emerged.

Two or more atoms that bond together form a(n)
Figure 1: Glow discharge in a low-pressure tube caused by electric current. Like what Faraday saw, the tube shows a dark space between the glows around the cathode (left, negatively charged) and anode (right, positively charged). image © Andrejdam/Wikimedia

Comprehension Checkpoint

Faraday is famous for discovering and naming the electron.

In 1857 the German glassblower Heinrich Geissler, while working for fellow countryman and physicist Julius Plücker at the University of Bonn, improved the quality of the vacuum that could be achieved in such tubes. However, Geissler’s tubes still contained enough gaseous atoms that when the electrical current travelled in the tube, there was an interaction between the two, causing the tubes to glow. In the mid-19th century these Geissler tubes were largely nothing more than a curiosity, but interestingly, a curiosity that proved to be a forerunner of neon lights.

Englishman William Crookes repeated experiments similar to those of Faraday and Geissler, but this time with the ‘new and improved’ vacuums. The number of gas atoms (and hence the pressure) was drastically reduced in Crookes’ tubes. This caused an interesting effect: Faraday’s dark space was observed even further down the tube, again extending away from cathode toward the anode.

In addition to the extension of the dark space, fluorescence was observed on the glass behind the anode at the positive end of the tube. When further experimentation revealed that shadows of objects placed in the tube were cast onto the glass behind the anode, the German physicist Johann Hittorf proposed that the shadows must have been created by something travelling in a straight line from the cathode to the anode. Yet another German physicist Eugen Goldstein, christened these invisible beams cathode rays.

As discussed in our module Early Ideas about Matter, Dalton’s atomic theory suggested that the atom was indivisible, i.e., that it was the smallest particle that made up matter, and that all matter was based upon that single unit. Experiments with cathode ray tubes dramatically changed that view when they led to the discovery of the first subatomic particle.

J.J. Thomson was an English physicist who worked with cathode ray tubes similar to those used by Crookes and others in the mid-19th century. Thomson’s experiments (Thomson, 1897) went further than those before him and provided evidence of the properties of the “something” hinted at by Hittorf. Thomson noted that cathode rays were deflected by magnetic fields and that the deflection was the same no matter what the source of the rays. This suggested that the rays were universal in their properties and that they had some magnetic charge. Thomson further demonstrated that cathode rays were charged because they could be deflected by an electric field. He found that the rays were deflected toward a positive plate and away from a negative plate, thus determining that they were made up of negatively charged particles of some sort.

Finally, Thomson applied both electrical and magnetic fields to the cathode ray at the same time. Knowing the strength of the fields applied and measuring the deflection of the particle stream in the tube, he was first able to measure the velocity of the particles in the stream. Then, by measuring the deflection of the ray while varying the two fields, Thomson was able to measure the mass-to-charge ratio of the particles in the stream and he found something astonishing. The negative particles had a mass-to-charge ratio that was over 1,000 times lower than that of a hydrogen atom, suggesting that the particles were incredibly tiny – much smaller than the smallest atom known. This fact allowed Thomson to definitively say that the atom was not the fundamental building block of matter and that smaller (subatomic) particles existed. Thomson originally called these particles corpuscles, but later they became known as electrons.

Comprehension Checkpoint

J.J. Thomson determined that cathode rays were made up of

Thomson’s discovery made sense of all of the previous observations made by Faraday, Geissler, and Crookes. Zipping through a tube filled with gas but partially under vacuum, electrons would eventually slam into those gas atoms, knocking off some of their electrons and making them fluoresce. The dark space that Faraday first noted was due to the distance needed for the electrons to accelerate to the speed necessary to ionize the tube’s gas atoms.

In the better vacuums achieved in the Crookes’ tubes, the electrons could travel further distances without interacting with gas molecules because of the lower density of molecules in the tube, thus extending the dark space.

Comprehension Checkpoint

The less gas there was in a cathode ray tube, the _____ dark space could be observed.

With the electron now discovered, Thomson went on to propose an entirely new model of the atom that was known as "The Plum Pudding Model." The model was so called since it mimicked the British desert of the same name that had dried fruit (primarily raisins not plums), dispersed in a body of suet and eggs that made a dough.

In his model Thomson proposed that the negatively charged electrons (analogous to the raisins) were randomly spread out among what he called "a sphere of uniform positive electrification" (analogous with the dough or body of the pudding) (see Figure 2).

Two or more atoms that bond together form a(n)
Figure 2: Thomson's "plum pudding model" of the atom, showing a positively-charged sphere containing many negatively-charged electrons in a random arrangement.

Thompson’s model of the atom as a doughy clump of positive and negative particles persisted until 1911, when Ernest Rutherford, a former student of Thomson’s, advanced atomic theory yet another notch.

During the years 1908–1911, Ernest Marsden and Hans Geiger performed a series of experiments under the direction of Ernest Rutherford at the University of Manchester in England. In these experiments alpha particles (tiny, positively charged particles) were fired at a thin piece of gold foil (Figure 3). Under the Thomson Plum Pudding model of the atom, the "sphere of uniform positive electrification" was thought to be so diffuse that the tiny, fast moving alpha particles would pass straight through. Similarly, the electrons in the model were thought to be so tiny that any electrostatic interactions between them and the positive alpha particles would be minimal, so the path of the alpha particles would hardly be affected.

Two or more atoms that bond together form a(n)
Figure 3: The gold foil experiment designed by Rutherford, Marsden, and Geiger. A beam of positively charged alpha particles was shot at a piece of gold foil. A screen around the foil captured the impact of the alpha particles.

As predicted, Rutherford and his co-workers observed that most of the alpha particles passed straight through the gold foil, and some particles were deflected at small angles. However, contradictory to what the Plum Pudding model predicted, a few rebounded at very sharp angles, some even flying straight back toward the source! These particles were acting as if they were encountering a hard object, like a tennis ball bouncing off a brick wall (Figure 4).

Two or more atoms that bond together form a(n)
Figure 4: In the gold foil experiment, Rutherford and his colleagues expected to see the alpha particles passing through the mostly empty "Plum Pudding"-style atoms. However, what they observed was that the alpha particles occasionally ricocheted at sharp angles, indicating there was something more solid in the atom than previously thought.

The fact that most of the alpha particles passed straight through the gold foil suggested to Rutherford that atoms are made up of largely empty space. However, contrary to the Thomson Plum Pudding model, Rutherford’s work suggested that there was a dense, positively charged area in an atom that caused the observed repulsion and backscattering of alpha particles. Rutherford was astonished by these observations and famously said:

It was quite the most incredible event that has ever happened to me in my life. It was almost as incredible as if you fired a 15-inch shell at a piece of tissue paper and it came back and hit you. On consideration, I realized that this scattering backward must be the result of a single collision, and when I made calculations I saw that it was impossible to get anything of that order of magnitude unless you took a system in which the greater part of the mass of the atom was concentrated in a minute nucleus. It was then that I had the idea of an atom with a minute massive centre, carrying a charge.

Over a series of experiments and papers (Rutherford, 1911, 1913, 1914), Rutherford developed a model of the atom with a dense, positively charged area of the atom at the center, now known as the nucleus – and the nuclear model of the atom was born.

Comprehension Checkpoint

Rutherford and his colleagues were surprised that

Following the discovery of the electron, Nobel Prize-winning physicist Robert Millikan conducted an ingenious experiment that allowed for the specific value of the negative charge of the electron to be calculated. In his famous oil drop experiment, Millikan and co-workers sprayed tiny oil droplets from an atomizer into a sealed chamber (Millikan, 1913). The oil drops fell downward, under the influence of gravity, into a space between two electrical plates. There they became charged, by interacting with air that had been ionized by X-rays.

Two or more atoms that bond together form a(n)
Figure 5: Millikan's oil drop experiment in which he observed droplets of oil fall between two electrical plates, where the droplets became ionized by X-rays.

By adjusting the voltage between the two electrical plates, Millikan applied an electrical force upward that exactly matched the gravitational force downward, thus suspending the drops motionless. When suspended, the electrical force and the force of gravity were working in opposite directions but were equal in magnitude. Hence:

where q is the charge on the oil drop, E is the electric field, m is the mass of the oil drop, and g is the gravitational field. By measuring the mass of each oil drop and knowing both the gravitational and the electrical field, the charge on each drop could be determined.

Millikan found that there were differing charges on different oil drops. However, in each case the charges on the oil drops were found to be multiples of 1.60 x 10-19 coulombs. He concluded that the differing charges were due to different numbers of electrons, each having a negative charge of 1.60 x 10-19 coulombs, and hence the charge on the electron was found.

Comprehension Checkpoint

Millikan found that the different oil drops in his experiment

Thomson’s electron and Rutherford’s nuclear model were tremendous advancements. The Japanese scientist Hantaro Nagaoka had previously rejected Thomson’s Plum Pudding model on the grounds that opposing charges could not penetrate each other, and he counter-proposed a model of the atom that resembled the planet Saturn with rings of electrons revolving around a positive center. Upon hearing of Rutherford’s work, he wrote to him in 1911 saying, "Congratulations on the simpleness of the apparatus you employ and the brilliant results you obtained."

But, the planetary model was not perfect, and several inconsistent experimental observations meant much work was still to be done. At the time the electron was still thought of as a small particle, and it was thought to spin almost randomly around the nucleus of the atom. It would take the additional experiments and the genius of Neils Bohr, Max Planck, and others to make the paradigm shift from classical physics in which atoms consist of tiny particles and are governed by laws of motion, to quantum mechanics in which electrons behave like waves and exhibit strange and exotic behaviors. (See our interactive animation comparing orbital and quantum models of the first 12 elements.) To learn more about the strange behaviors of quantum physics, read the other entries in our Atomic Theory series: II: Bohr and the Beginnings of Quantum Theory, III: Wave-Particle Duality and the Electron, and IV: Quantum Numbers and Orbitals.

Two or more atoms that bond together form a(n)

Interactive Animation: Atomic and ionic structure of the first 12 elements

The 19th and early 20th centuries saw great advances in our understanding of the atom. This module takes readers through experiments with cathode ray tubes that led to the discovery of the first subatomic particle: the electron. The module then describes Thomson’s plum pudding model of the atom along with Rutherford’s gold foil experiment that resulted in the nuclear model of the atom. Also explained is Millikan’s oil drop experiment, which allowed him to determine an electron’s charge. Readers will see how the work of many scientists was critical in this period of rapid development in atomic theory.

Key Concepts

  • Atoms are not dense spheres but consist of smaller particles including the negatively charged electron.

  • The research on passing electrical currents through vacuum tubes by Faraday, Geissler, Crookes, and others laid the groundwork for discovery of the first subatomic particle.

  • J.J. Thomson’s observations of cathode rays provide the basis for the discovery of the electron.

  • Rutherford, Geiger, and Marsden performed a series of gold foil experiments that indicated that atoms have small, dense, positively-charged centers – later named the nucleus.

  • Millikan’s oil drop experiment determines the fundamental charge on the electron as 1.60 x 10-19 coulombs.

  • HS-C4.4, HS-C6.2, HS-PS1.A1, HS-PS1.A3
  • Faraday, M. (1838). VIII. Experimental researches in electricity. Thirteenth series. Philosophical Transactions of the Royal Society of London, 128: 125-168.

  • Millikan, R.A. (1913). On the elementary electric charge and the Avogadro Constant. Physics Review, 2(2): 109–143.
  • Rutherford, E. (1911). The scattering of α and β particles by matter and the structure of the atom. Philosophical Magazine, Series 6, 21(125): 669–688.
  • Rutherford, E., & Nuttal, J.M. (1913). Scattering of α-particles by gases. Philosophical Magazine, Series 6, 26(154): 702–712.
  • Rutherford, E. (1914). The structure of the atom. Philosophical Magazine. Series 6, 27(159): 488–498.
  • Thomson, J.J. (1897). Cathode rays. Philosophical Magazine, Series 5, 44(269): 293-316.

Adrian Dingle, B.Sc., Anthony Carpi, Ph.D. “Atomic Theory I” Visionlearning Vol. CHE-1 (2), 2003.

Top


Page 10

Atomic Theory and Structure

by Adrian Dingle, B.Sc., Anthony Carpi, Ph.D.

The earliest ideas about matter at the atomic level were built over many centuries. Starting with the ancient Greeks, and moving through to the beginning of the 19th century, the story unfolds relatively slowly. (You can read more about this is in our modules Early Ideas about Matter: From Democritus to Dalton and Atomic Theory I: The Early Days.) Despite the slow pace, it is crucial to understand that the process was a methodical one as each scientist built upon earlier ideas. This gradual, logical progression, where atomic structure evolved from being a simple, philosophical idea, through to the ultra-sophisticated world of the Higgs boson particle discovered in the early part of the 21st century, represents a wonderful example of the evolution of a scientific idea, and the application of the scientific process. In fact, one could argue that the history, struggle, and achievement that is threaded through the development of understanding matter at the atomic level is the quintessential story of the scientific method.

The story of atomic theory first encounters reproducible, scientific (evidence based) proof in the late 18th century. French chemists Antoine Lavoisier and Joseph Proust, with their Law of Conservation of Mass in 1789 and Law of Definite Proportions in 1799, respectively, each laid the groundwork for Englishman John Dalton’s work on the Law of Multiple Proportions (Dalton, 1803). Given that many centuries had elapsed between the earliest ideas of the atom and Dalton’s work, it would be fair to say that the evolution of atomic theory had been a gradual one, with progression in the field being steady rather than spectacular. But that was all about to change, and quite dramatically.

The most intense period of progress took place between the late 19th and early 20th century, and it hinged heavily on the work of a Danish physicist named Niels Bohr. Like so many before him, Bohr built upon the work of his predecessors, and for Bohr, part of that foundation had been built by Ernest Rutherford.

Based upon a series of experiments, Rutherford proposed the planetary model of the atom in which electrons swirled around a hard, dense nucleus (see Atomic Theory I: The Early Days). While Rutherford’s model explained many observations accurately, it was found to have flaws.

Rutherford’s planetary model of the atom was based upon classical physics – a system that deals with physical particles, force, and momentum. Unfortunately, this same system predicted that electrons orbiting in the manner that Rutherford described would lose energy, give off radiation, and ultimately crash into the nucleus and destroy the atom. However, for the most part, atoms are stable, lasting literally billions of years. Furthermore, the radiation predicted by the Rutherford model would have been a continuous spectrum of every color – in essence white light that when passed through a prism would display all of the colors of the rainbow (Figure 1).

Two or more atoms that bond together form a(n)
Figure 1: When passed through a prism, white light displays the color spectrum.

But when pure gases of different elements are excited by electricity, as they would have been when placed in the newly discovered electric-discharge tube, they emit radiation at distinct frequencies. In other words, different elements do not emit white light, they emit light of different colors, and when that light is passed through a prism it does not produce a continuous rainbow of colors, but a pattern of colored lines, now referred to as line spectra (Figure 2). Clearly, Rutherford’s model did not fit with all of the observations, and Bohr made it his business to address these inconsistencies.

Two or more atoms that bond together form a(n)
Figure 2: The visible light spectrum is displayed at the top and line spectra for three elements - hydrogen, neon, and iron - are below. image © Neon spectrum: Deo Favente

In 1911, Niels Bohr (Figure 3) had just completed his doctorate in physics at the University of Copenhagen and was invited to continue his work at the University of Manchester in England by Rutherford. Rutherford had already significantly advanced atomic theory with his groundbreaking gold foil experiment, but Bohr’s genius was in taking the Rutherford model and advancing it further.

Two or more atoms that bond together form a(n)
Figure 3: Niels Bohr

It wasn’t Bohr who had come up with the original idea of a planetary model of the atom, but he was the one who took the fundamental concept and applied new ideas about quantum theory to it. This leap was necessary to explain the new evidence that challenged the old model, and to subsequently formulate a new, ‘better’ model.

Two or more atoms that bond together form a(n)

Interactive Animation: Bohr’s Atom

Comprehension Checkpoint

The planetary model of the atom was based on

It’s easy to think of light, and other forms of energy, as continuous. Turn up the dimmer switch on your lamp, and the lamp gets gradually brighter. However, by the late 1800s, physicists were beginning to suspect that this was not in fact true. Classical physics models failed to accurately predict black-body radiation; in other words, classical physics did not accurately predict the energy given off by an object when it was heated.

The German physicist Max Planck solved this problem in 1903 by proposing that black-body radiation energy had to be quantized, i.e., that it could only be released or absorbed in specific ‘packets’ that were associated with specific frequencies. This solved the black-body problem and was consistent with the observed experimental data. Thus, quantum mechanics was born.

Despite the advances that he and others made using this idea, interestingly, Planck remained quite skeptical of quantized energy for many years. He insisted that the calculations that he had done, and the conclusions that he had reached, were somehow a sophisticated mathematical trick and that ultimately the old, classical model would prevail. After all, it had been around for approximately 200 years and had stood up to some pretty intense scrutiny.

In 1905, Albert Einstein published a series of papers proposing that light also exhibited quantum behavior (Einstein, 1905). Sometimes described as Einstein’s Annus mirabilis (miracle year), the papers taken together, and combined with Planck’s work, allowed Bohr to marry the nature of the atom with physics to usher in a new dawn of understanding in atomic theory.

In 1913, building on Planck’s and Einstein’s theories of quantization, Bohr proposed that the electron itself was quantized – that it could not exist just anywhere around an atom (as suggested by the Rutherford model) but instead could only be found in specific positions, with specific energies. The electron could transition to different positions, but only in discrete, defined steps. It could not spin in any location around the nucleus of an atom but instead was restricted to specific areas of space – much like the planets in our solar system are restricted to specific paths.

Being negatively charged, electrons are attracted to the positive protons in the nucleus of an atom, and will normally occupy the orbital, or path, within an atom that is closest to the nucleus if it is available. This state, which has low potential energy, is called the ground state. By exposing the electrons to an external source of energy such as an electric discharge, it is possible to promote the electrons from their ground state to other positions that have higher potential energies, called excited states. These ‘excited’ electrons quickly return to lower energy positions (in order to regain the stability associated with lower energies), and in doing so they release energy in specific frequencies that correspond to the energy differences between electron orbitals, or shells (see quantum behavior simulation). Bohr’s mathematical equations further predicted that electrons would not crash into the nucleus in a manner that classical physics – and Rutherford’s model – had predicted. This was another crucial realization that made a jump from one paradigm (classical physics) to a new one (quantum physics).

Two or more atoms that bond together form a(n)

Interactive Animation: Atomic and ionic structure of the first 12 elements

Bohr’s discovery that Planck’s quantum theory could be applied to the classical Rutherford model of the atom, and could account for the observed shortcomings in the original model, is another beautiful example of how scientific theory uses prior evidence, coupled with new experimental observations, to adapt, develop, and change models and understanding over time. Science is usually advanced by contemporary scientists building on the work of predecessors and, as Isaac Newton put it in his 1676 letter to Robert Hooke (both prominent scientists of their time), by their “standing on the shoulders of giants.” Bohr’s work built on the theories of those before him, and extended them to explain the experimentally observed line spectra of atoms in a mathematical proof that made perfect sense.

Comprehension Checkpoint

In an atom, the ground state is the orbital with the ______ potential energy.

While Bohr’s work seemed to explain the curious phenomenon of line spectra, a number of lines had been observed in the spectra of hydrogen that did not fit Bohr’s theory. At first glance this appeared to poke holes in Bohr’s ideas, but Bohr was quick to offer an explanation. He suggested that the lines in the spectrum for hydrogen that could not be accounted for were actually caused not by hydrogen atoms, but rather by an entirely different species altogether. So, what were these different species, and how did they come to be?

Almost 30 years prior to Bohr's publishing his famous trilogy of papers in the Philosophical Magazine and Journal of Science in 1913, the idea of particles having the ability to carry some kind of charge had been established by the Swedish scientist Svante Arrhenius and the Englishman Michael Faraday. These charged particles had been christened ions.

Atoms are electrically neutral, meaning that the number of positive protons in any given atom equals the number of negative electrons in the same. Since the positive charge is exactly cancelled by the negative, the atom has no overall electrical charge. However, just as it is possible to excite an electron to a higher orbital, as Bohr’s work had shown, it is also possible to give an electron sufficient energy to overcome the attraction of the nucleus completely and to remove it from the atom entirely. This has the effect of unbalancing the electrical charge and results in the formation of a species with an overall positive charge – called a cation. For example, a sodium atom can lose an electron to form a positively charged sodium cation (Equation 1); the energy associated with the ejection of the first electron from any atom is called the first ionization energy.

Na(s) → Na+(s) + e-

Cations that are formed by the ejection of one electron can be further ionized by losing additional electrons, and in the process forming another ion, this time with a 2+ charge. The energy required for the ejection of the second electron is known as the second ionization energy. Although it rarely occurs with larger atoms (or indeed with the sodium example given above), it is theoretically possible to remove all of the electrons from any given atom, leading to third, fourth, fifth, etc. ionization energies for atoms with large numbers of electrons.

Two or more atoms that bond together form a(n)
Figure 4: Using the element hydrogen, examples of a cation and anion. image © Jkwchui

Once an electron (or electrons) has been ejected from an atom in this manner, it can be accepted by other atoms, and as such, electrons can be transferred from one atom to another. Just as releasing electrons unbalances the charge, accepting electrons causes atoms to become unbalanced in terms of their charge as well, and once again an ion is formed. This time, the ion has a negative charge (a greater number of electrons than protons), and the species is called an anion. (A hydrogen cation and anion are shown in Figure 4.) For example, a neutral chlorine atom, with equal numbers of protons and electrons, can accept an electron from an external source to form a negatively charged chloride anion (Equation 2).

Cl(g) + e- → Cl-(g)

Ions have wildly different properties when compared to their parent atoms, and this transfer of electrons from one atom to another is an example of how a small change in the structure of an atom can make a large difference in the behavior and nature of the particles. For example, sodium metal is unstable, reacting violently with water and corroding instantaneously in air. Thus, coming into contact with free sodium metal would be extremely dangerous. Similarly, chlorine exists as a gas under ambient conditions and it is highly poisonous, scarring the lungs of anyone who breathes it (in fact, free chlorine gas was used as a chemical weapon in World War I). However, when the two substances react with one another, sodium loses an electron forming a cation, and chlorine accepts the same electron to form an anion. The two resulting ions then bond together as a result of their charges, and together create a very common substance – table salt, which is neither reactive nor poisonous.

Comprehension Checkpoint

_____ have a positive charge, while _____ have a negative charge.

In Bohr’s work, the ionization of one particular element, helium, proved to be the key to unlock the explanation for the unexpected lines that he observed in hydrogen’s spectrum. When a helium atom, which has two electrons, loses an electron to form the helium ion, He+, its electronic structure mimics that of atomic hydrogen, since both species only have only one electron – the helium ion is said to be isoelectronic with the hydrogen atom. However, the helium ion possesses a nucleus with double the charge of a hydrogen atom (two protons as opposed to one proton). Bohr realized this, and suggested that the greater attraction between the electron and He nucleus accounted for the spectral lines that were previously unexplained – the charge of the nucleus affected the energy associated with transfers of electrons between the orbitals. Bohr’s theory was proven correct when spectra were generated using ionized helium that had been purged of hydrogen.

At the close of the 19th century, two different particles were known to exist in the atom, and both possessed an electrical charge – the very small and negatively charged electron and the much larger and positively charged proton. However, by the beginning of the 20th century, evidence began to mount that this was not a complete picture of the atom. Specifically, the mass of protons and electrons in an atom did not appear sufficient to justify the mass of the whole atom, and certain types of nuclear decay suggested that something else might be going on in the nucleus. In 1932, James Chadwick, a British physicist who had studied with, and was working for, Ernest Rutherford at the time, set out to solve the problem. Rutherford had proposed the idea of a neutral atomic particle that had mass as early as 1920, but he never managed to gain traction in the hunt for this mysterious particle.

In 1932 Chadwick further developed an experiment that had been first performed by Frederic Joliot-Curie and Irene Joliot-Curie. They found that by using polonium as a source of alpha particles, they could cause beryllium to emit radiation that, in turn, could be used to knock protons out of a piece of paraffin wax. The Joliot-Curies proposed that this radiation was gamma radiation, a packet of energy with no true mass. As an accomplished researcher of gamma rays and of the nucleus of the atom, Chadwick realized what others had not – that protons were too massive to be ejected from paraffin by mass-less gamma rays. By more carefully measuring the impact of the mystery particle on the paraffin wax and combining this with other measurements, Chadwick concluded that the particles being emitted were not gamma radiation, but a relatively heavy particle that had no charge – a particle named the neutron (Figure 5).

Two or more atoms that bond together form a(n)
Figure 5: Artistic model of an atom showing the nucleus, with protons and neutrons, and orbiting electrons. image © Visionlearning

Chadwick wrote a paper about his discovery entitled “The Possible Existence of a Neutron,” and it was published in the journal Nature (Chadwick, 1932). In 1935, he was awarded the Nobel Prize in Physics for his discovery. The Joliot-Curies did not go without recognition either. Their work on radioactivity and radioactive isotopes won them the Nobel Prize for Chemistry, also in 1935.

Chadwick’s discovery marked the genesis of induced nuclear reactions, where the neutron is accelerated and crashed into the nuclei of other elements, generating massive amounts of energy (the neutron can easily do this since being neutral it is not repelled from the nucleus in the way that positively charged particles would be). These reactions had a massive impact upon the world as a whole since they spawned thoughts of the atomic bomb and of nuclear energy.

Comprehension Checkpoint

A _____ is a heavy particle that has no charge.

The neutron also explains the existence of atoms of the same element that have different atomic masses. Isotopes are atoms of the same element (i.e., they have the same numbers of protons) but that differ in the number of neutrons they possess. As a result, different isotopes have similar chemical properties, but their masses, and in some cases their physical behavior, differ. Isotopes are differentiated by their atomic mass, which can be indicated by writing the element’s symbol, followed by a dash and then the mass, or, more commonly, by writing the mass as a superscript before the element symbol. For example, carbon-12 (C-12, or 12C) and carbon-14 (C-14, or 14C) are both naturally occurring isotopes of carbon. Carbon-12 is a stable isotope that accounts for almost 99% of naturally occurring carbon. Carbon-14 is a radioactive isotope that accounts for only about 1 x 10-10 % of naturally occurring carbon, but because it decays to nitrogen with a half-life of approximately 5,730 years, it can be used to date some carbon-containing objects. (Three isotopes of carbon are depicted in Figure 6.)

Two or more atoms that bond together form a(n)
Figure 6: Carbon isotopes. Each have the same number of protons, but different numbers of neutrons.

There is often more than one naturally occurring isotope of any given, individual element. As a result, the atomic masses that are given on a modern periodic table are the weighted masses of all the known isotopes of each element. For example, naturally occurring chlorine has two principal isotopes, one with 18 neutrons (mass of 35), and one with 20 neutrons (mass of 37). The lighter isotope 35Cl, accounts for almost 76% of its natural abundance, while the 37Cl isotope accounts for only about 24%. Thus the weighted mean is the average mass of naturally occurring chlorine atoms weighted by their relative abundance, or 35.45.

Comprehension Checkpoint

Isotopes are atoms of the same element that have a different

Bohr’s work provided a bridge between several disparate ideas that may never have been linked together without his intervention. He changed the paradigm from applying classical (particle) physics to the model of the atom, to thinking about the application of quantum theory and waves – a truly crucial development in the grand scheme of atomic theory and one that laid the groundwork for future scientists to build upon.

Having traveled from the earliest work on atomic theory through the crucial early part of the 20th century, we still have some way to go in the story of the atom. The advancements described in this module were grounded in Rutherford’s work, modified by Planck’s insights, and pieced together by Bohr’s genius. Still, there would be at least another decade that had to pass before work by Pauli, Heisenberg, and ultimately Schrödinger led to the full development of modern quantum mechanics that completely describes the atom as we know it today.

This module is an updated version of our previous content, to see the older module please go to this link.

The 20th century brought a major shift in our understanding of the atom, from the planetary model that Ernest Rutherford proposed to Niels Bohr’s application of quantum theory and waves to the behavior of electrons. With a focus on Bohr’s work, the developments explored in this module were based on the advancements of many scientists over time and laid the groundwork for future scientists to build upon further. The module also describes James Chadwick’s discovery of the neutron. Among other topics are anions, cations, and isotopes.

Key Concepts

  • Drawing on experimental and theoretical evidence, Niels Bohr changed the paradigm of modern atomic theory from one that was based on physical particles and classical physics, to one based in quantum principles.

  • Under Bohr’s model of the atom, electrons cannot rotate freely around the atom, but are bound to certain atomic orbitals that both constrain and define an atom's electronic behavior.

  • Atoms can gain or lose electrons to become electrically charged ions.

  • James Chadwick completed the early picture of the atom with his discovery of the neutron, a neutral, nuclear particle that affects an atom’s mass and the different physical properties of atomic isotopes.

  • HS-C4.4, HS-C6.2, HS-PS1.A1, HS-PS1.A3
  • Bohr, N. (1913). On the Constitution of Atoms and Molecules. Philosophical Magazine (London), Series 6 (26), 1–25.

  • Chadwick, J. (1932). The Possible Existence of a Neutron. Nature, 129(3252), 312.
  • Dalton, John (1805). On the Absorption of Gases by Water and Other Liquids. Memoirs of the Literary and Philosophical Society of Manchester, Series 2(1), 271–287.
  • Einstein, A. (1905). A New Determination of Molecular Dimensions. Annalen der Physik, Series 4(19), 289–306.
  • Einstein, A. (1905). Does the inertia of a body depend on its energy content? Annalen der Physik, Series 4(18), 639–641.
  • Einstein, A. (1905). On the electrodynamics of moving bodies. Annalen der Physik, Series 4(17), 891–921.
  • Einstein, A. (1905). On a heuristic viewpoint concerning the production and transformation of light. Annalen der Physik, Series 4(17), 132–148.
  • Einstein, A. (1905). On the motion of small particles suspended in liquids at rest required by the molecular-kinetic theory of heat. Annalen der Physik, Series 4(19), 371–381.
  • Planck, M. (1903). Treatise on Thermodynamics. Ogg, A. (trans.). London: Longmans, Green & Co.

Adrian Dingle, B.Sc., Anthony Carpi, Ph.D. “Atomic Theory II” Visionlearning Vol. CHE-1 (3), 2003.

Top


Page 11

Atomic Theory and Structure

by Adrian Dingle, B.Sc., Anthony Carpi, Ph.D.

As discussed in our Atomic Theory II module, at the end of 1913 Niels Bohr facilitated the leap to a new paradigm of atomic theory – quantum mechanics. Bohr’s new idea that electrons could only be found in specified, quantized orbits was revolutionary (Bohr, 1913). As is consistent with all new scientific discoveries, a fresh way of thinking about the universe at the atomic level would only lead to more questions, the need for additional experimentation and collection of evidence, and the development of expanded theories. As such, at the beginning of the second decade of the 20th century, another rich vein of scientific work was about to be mined.

In the late 19th century, the father of the periodic table, Russian chemist Dmitri Mendeleev, had already determined that the elements could be grouped together in a manner that showed gradual changes in their observed properties. (This is discussed in more detail in our module The Periodic Table of Elements.) By the early 1920s, other periodic trends, such as atomic volume and ionization energy, were also well established.

Two or more atoms that bond together form a(n)
The Periodic Table of Elements

The German physicist Wolfgang Pauli made a quantum leap by realizing that in order for there to be differences in ionization energies and atomic volumes among atoms with many electrons, there had to be a way that the electrons were not all placed in the lowest energy levels. If multi-electron atoms did have all of their electrons placed in the lowest energy levels, then very different periodic patterns would have resulted from what was actually observed. However, before we reach Pauli and his work, we need to establish a number of more fundamental ideas.

The development of early quantum theory leaned heavily on the concept of wave-particle duality. This simultaneously simple and complex idea is that light (as well as other particles) has properties that are consistent with both waves and particles. The idea had been first seriously hinted at in relation to light in the late 17th century. Two camps formed over the nature of light: one in favor of light as a particle and one in favor of light as a wave. (See our Light I: Particle or Wave? module for more details.) Although both groups presented effective arguments supported by data, it wasn’t until some two hundred years later that the debate was settled.

Two or more atoms that bond together form a(n)

Interactive Animation: Atomic and ionic structure of the first 12 elements

At the end of the 19th century the wave-particle debate continued. James Clerk Maxwell, a Scottish physicist, developed a series of equations that accurately described the behavior of light as an electromagnetic wave, seemingly tipping the debate in favor of waves. However, at the beginning of the 20th century, both Max Planck and Albert Einstein conceived of experiments which demonstrated that light exhibited behavior that was consistent with it being a particle. In fact, they developed theories that suggested that light was a wave-particle – a hybrid of the two properties. By the time of Bohr’s watershed papers, the time was right for the expansion of this new idea of wave–particle duality in the context of quantum theory, and in stepped French physicist Louis de Broglie.

In 1924, de Broglie published his PhD thesis (de Broglie, 1924). He proposed the extension of the wave-particle duality of light to all matter, but in particular to electrons. The starting point for de Broglie was Einstein’s equation that described the dual nature of photons, and he used an analogy, backed up by mathematics, to derive an equation that came to be known as the “de Broglie wavelength” (see Figure 1 for a visual representation of the wavelength).

The de Broglie wavelength equation is, in the grand scheme of things, a profoundly simple one that relates two variables and a constant: momentum, wavelength, and Planck's constant. There was support for de Broglie’s idea since it made theoretical sense, but the very nature of science demands that good ideas be tested and ultimately demonstrated by experiment. Unfortunately, de Broglie did not have any experimental data, so his idea remained unconfirmed for a number of years.

Two or more atoms that bond together form a(n)
Figure 1: Two representations of a de Broglie wavelength (the blue line) using a hydrogen atom: a radial view (A) and a 3D view (B).

It wasn’t until 1927 that de Broglie’s hypothesis was demonstrated via the Davisson-Germer experiment (Davisson, 1928). In their experiment, Clinton Davisson and Lester Germer fired electrons at a piece of nickel metal and collected data on the diffraction patterns observed (Figure 2). The diffraction pattern of the electrons was entirely consistent with the pattern already measured for X-rays and, since X-rays were known to be electromagnetic radiation (i.e., waves), the experiment confirmed that electrons had a wave component. This confirmation meant that de Broglie’s hypothesis was correct.

Two or more atoms that bond together form a(n)
Figure 2: A drawing of the experiment conducted by Davisson and Germer where they fired electrons at a piece of nickel metal and observed the diffraction patterns. image © Roshan220195

Interestingly, it was the (experimental) efforts of others (Davisson and Germer), that led to de Broglie winning the Nobel Prize in Physics in 1929 for his theoretical discovery of the wave-nature of electrons. Without the proof that the Davisson-Germer experiment provided, de Broglie’s 1924 hypothesis would have remained just that – a hypothesis. This sequence of events is a quintessential example of a theory being corroborated by experimental data.

Comprehension Checkpoint

Theories must be backed up by

In 1926, Erwin Schrödinger derived his now famous equation (Schrödinger, 1926). For approximately 200 years prior to Schrödinger’s work, the infinitely simpler F = ma (Newton’s second law) had been used to describe the motion of particles in classical mechanics. With the advent of quantum mechanics, a completely new equation was required to describe the properties of subatomic particles. Since these particles were no longer thought of as classical particles but as particle-waves, Schrödinger’s partial differential equation was the answer. In the simplest terms, just as Newton’s second law describes how the motion of physical objects changes with changing conditions, the Schrödinger equation describes how the wave function (Ψ) of a quantum system changes over time (Equation 1). The Schrödinger equation was found to be consistent with the description of the electron as a wave, and to correctly predict the parameters of the energy levels of the hydrogen atom that Bohr had proposed.

Two or more atoms that bond together form a(n)
Equation 1: The Schrödinger equation.

Schrödinger’s equation is perhaps most commonly used to define a three-dimensional area of space where a given electron is most likely to be found. Each area of space is known as an atomic orbital and is characterized by a set of three quantum numbers. These numbers represent values that describe the coordinates of the atomic orbital: including its size (n, the principal quantum number), shape (l, the angular or azimuthal quantum number), and orientation in space (m, the magnetic quantum number). There is also a fourth quantum number that is exclusive to a particular electron rather than a particular orbital (s, the spin quantum number; see below for more information).

Schrödinger’s equation allows the calculation of each of these three quantum numbers. This equation was a critical piece in the quantum mechanics puzzle, since it brought quantum theory into sharp focus via what amounted to a mathematical demonstration of Bohr’s fundamental quantum idea. The Schrödinger wave equation is important since it bridges the gap between classical Newtonian physics (which breaks down at the atomic level) and quantum mechanics.

The Schrödinger equation is rightfully considered to be a monumental contribution to the advancement and understanding of quantum theory, but there are three additional considerations, detailed below, that must also be understood. Without these, we would have an incomplete picture of our non-relativistic understanding of electrons in atoms.

German mathematician and physicist Max Born made a very specific and crucially important contribution to quantum mechanics relating to the Schrödinger equation. Born took the wave functions that Schrödinger produced, and said that the solutions to the equation could be interpreted as three-dimensional probability “maps” of where an electron may most likely be found around an atom (Born, 1926). These maps have come to be known as the s, p, d, and f orbitals (Figure 3).

Two or more atoms that bond together form a(n)
Figure 3: Based on Born's theories, these are representations of the three-dimensional probabilities of an electron's location around an atom. The four orbitals, in increasing complexity, are: s, p, d, and f. Additional information is given about the orbital's magnetic quantum number (m). image © UC Davis/ChemWiki

In the year following the publication of Schrödinger’s work, the German physicist Werner Heisenberg published a paper that outlined his uncertainty principle (Heisenberg, 1927). He realized that there were limitations on the extent to which the momentum of an electron and its position could be described. The Heisenberg Uncertainty Principle places a limit on the accuracy of simultaneously knowing the position and momentum of a particle: As the certainty of one increases, then the uncertainty of other also increases.

The crucial thing about the uncertainty principle is that it fits with the quantum mechanical model in which electrons are not found in very specific, planetary-like orbits – the original Bohr model – and it also dovetails with Born’s probability maps. The two contributions (Born and Heisenberg’s) taken together with the solution to the Schrödinger equation, reveal that the position of the electron in an atom can only be accurately predicted in a statistical way. That is to say, we know where the electron is most likely to be found in the atom, but we can never be absolutely sure of its exact position.

Comprehension Checkpoint

The Heisenberg uncertainty principle concerning the position and momentum of a particle states that as the certainty of one increases, the _____ of the other increases.

In 1922 German physicists Otto Stern, an assistant of Born’s, and Walther Gerlach conducted an experiment in which they passed silver atoms through a magnetic field and observed the deflection pattern. In simple terms, the results yielded two distinct possibilities related to the single, 5s valence electron in each atom. This was an unexpected observation, and implied that a single electron could take on two, very distinct states. At the time, nobody could explain the phenomena that the experiment had demonstrated, and it took a number of scientists, working both independently and in unison with earlier experimental observations, to work it out over a period of several years.

In the early 1920s, Bohr’s quantum model and various spectra that had been produced could be adequately described by the use of only three quantum numbers. However, there were experimental observations that could not be explained via only three mathematical parameters. In particular, as far back as 1896, the Dutch physicist Pieter Zeeman noted that the single valence electron present in the sodium atom could yield two different spectral lines in the presence of a magnetic field. This same phenomenon was observed with other atoms with odd numbers of valence electrons. These observations were problematic since they failed to fit the working model.

In 1925, Dutch physicist George Uhlenbeck and his graduate student Samuel Goudsmit proposed that these odd observations could be explained if electrons possessed angular momentum, a concept that Wolfgang Pauli later called “spin.” As a result, the existence of a fourth quantum number was revealed, one that was independent of the orbital in which the electron resides, but unique to an individual electron.

By considering spin, the observations by Stern and Gerlach made sense. If an electron could be thought of as a rotating, electrically-charged body, it would create its own magnetic moment. If the electron had two different orientations (one right-handed and one left-handed), it would produce two different ‘spins,’ and these two different states would explain the anomalous behavior noted by Zeeman. This observation meant that there was a need for a fourth quantum number, ultimately known as the “spin quantum number,” to fully describe electrons. Later it was determined that the spin number was indeed needed, but for a different reason – either way, a fourth quantum number was required.

Comprehension Checkpoint

Some experimental observations could not be explained mathematically using three parameters because

In 1922, Niels Bohr visited his colleague Wolfgang Pauli at Göttingen where he was working. At the time, Bohr was still wrestling with the idea that there was something important about the number of electrons that were found in ‘closed shells’ (shells that had been filled).

In his own later account (1946), Pauli describes how building upon Bohr’s ideas and drawing inspiration from others’ work, he proposed the idea that only two electrons (with opposite spins) should be allowed in any one quantum state. He called this ‘two-valuedness’ – a somewhat inelegant translation of the German zweideutigkeit (Pauli, 1925). The consequence was that once a pair of electrons occupies a low energy quantum state (orbitals), any subsequent electrons would have to enter higher energy quantum states, also restricted to pairs at each level.

Using this idea, Bohr and Pauli were able to construct models of all of the electronic structures of the atoms from hydrogen to uranium, and they found that their predicted electronic structures matched the periodic trends that were known to exist from the periodic table – theory met experimental evidence once again.

Pauli ultimately formed what came to be known as the exclusion principle (1925), which used a fourth quantum number (introduced by others) to distinguish between the two electrons that make up the maximum number of electrons that could be in any given quantum level. In its simplest form, the Pauli exclusion principle states that no two electrons in an atom can have the same set of four quantum numbers. The first three quantum numbers for any two electrons can be the same (which places them in the same orbital), but the fourth number must be either +½ or -½, i.e., they must have different ‘spins’ (Figure 4). This is what Uhlenbeck and Goudsmit’s research suggested, following Pauli’s original publication of his theories.

Two or more atoms that bond together form a(n)
Figure 4: A model of the fourth quantum number, spin (s). Shown here are models for particles with spin (s) of ½, or half angular momentum.

The period described here was rich in the development of the quantum theory of atomic structure. Literally dozens of individuals, some mentioned throughout this module and others not, contributed to this process by providing theoretical insights or experimental results that helped shape our understanding of the atom. Many of the individuals worked in the same laboratories, collaborated together, or communicated with one another during the period, allowing the rapid transfer of ideas and refinements that would shape modern physics. All these contributions can certainly been seen as an incremental building process, where one idea leads to the next, each adding to the refinement of thinking and understanding, and advancing the science of the field.

The 20th century was a period rich in advancing our knowledge of quantum mechanics, shaping modern physics. Tracing developments during this time, this module covers ideas and refinements that built on Bohr’s groundbreaking work in quantum theory. Contributions by many scientists highlight how theoretical insights and experimental results revolutionized our understanding of the atom. Concepts include the Schrödinger equation, Born’s three-dimensional probability maps, the Heisenberg uncertainty principle, and electron spin.

Key Concepts

  • Electrons, like light, have been shown to be wave-particles, exhibiting the behavior of both waves and particles.

  • The Schrödinger equation describes how the wave function of a wave-particle changes with time in a similar fashion to the way Newton’s second law describes the motion of a classic particle. Using quantum numbers, one can write the wave function, and find a solution to the equation that helps to define the most likely position of an electron within an atom.

  • Max Born’s interpretation of the Schrödinger equation allows for the construction of three-dimensional probability maps of where electrons may be found around an atom. These ‘maps’ have come to be known as the s, p, d, and f orbitals.

  • The Heisenberg Uncertainty Principle establishes that an electron’s position and momentum cannot be precisely known together, instead we can only calculate statistical likelihood of an electron’s location.

  • The discovery of electron spin defines a fourth quantum number independent of the electron orbital but unique to an electron. The Pauli exclusion principle states that no two electrons with the same spin can occupy the same orbital.

  • HS-C1.4, HS-C4.4, HS-PS1.A2, HS-PS2.B3
  • Bohr, N. (1913). On the constitution of atoms and molecules. Philosophical Magazine (London), Series 6, 26, 1–25.

  • Born, M. (1926). Zur Quantenmechanik der Stoßvorgänge. Zeitschrift für Physik, 37(12), 863–867.
  • Davisson, C. J. (1928). Are electrons waves? Franklin Institute Journal, 205(5), 597-623.
  • de Broglie, L. (1924). Recherches sur la théorie des quanta. Annales de Physique, 10(3), 22-128.
  • Heisenberg, W. (1927). Über den anschaulichen Inhalt der quantentheoretischen Kinematik und Mechanik. Zeitschrift für Physik, 43(3-4), 172-198.
  • Pauli, W. (1925). Ueber den Einfluss der Geschwindigkeitsabhaengigkeit der Elektronenmasse auf den Zeeman-Effekt. Zeitschrift für Physik, 31(1), 373-385.
  • Pauli, W. (1946). Remarks on the history of the exclusion principle. Science, New Series, 103(2669), 213-215.
  • Schrödinger, E. (1926). Quantisierung als Eigenwertproblem. Annalen der Physik, 384(4), 273–376.
  • Stoner, E. C. (1924). The distribution of electrons among atomic energy levels. The London, Edinburgh and Dublin Philosophical Magazine (6th series), 48(286), 719-736

Adrian Dingle, B.Sc., Anthony Carpi, Ph.D. “Atomic Theory III” Visionlearning Vol. CHE-3 (6), 2015.

Top


Page 12

Atomic Theory and Structure

by Adrian Dingle, B.Sc., Anthony Carpi, Ph.D.

In Atomic Theory III: Wave-Particle Duality and the Electron, we discussed the advances that were made by Schrödinger, Born, Pauli, and others in the application of the quantum model to atomic theory. The Schrödinger equation was seen as a key mathematical link between the theory and the application of the quantum model. Born took the wave functions that Schrödinger produced and said that the solutions to the equation could define the energies and the most probable positions of electrons within atoms, thus allowing us to build a much more detailed description of where electrons might be found within an atom. This module further explores these solutions, the position of electrons, the shape of atomic orbitals, and the implications of these ideas.

As we saw in earlier reading, the electron is not a true particle, but a wave-particle similar to the photon. Since we measure the position of objects with light, the small size of the electron introduces a challenge. If we shine a light beam on a moving tennis ball, the light has little effect on the tennis ball and we can measure both its position and momentum with a high degree of accuracy. However, the electron is so tiny that even a single photon will influence its trajectory – thus if we shine a beam of light on it to measure its position, the energy of the photon will affect its momentum, and vice versa. Werner Heisenberg developed a principle to describe this uncertainty, called appropriately the Heisenberg Uncertainty Principle (Heisenberg, 1927). It tells us that mathematically, the product of the uncertainty in position (Δx) and the uncertainty in momentum (Δp) of an electron cannot be less than the reduced Planck constant ℏ/2 (Equation 1).

Two or more atoms that bond together form a(n)
Equation 1: The Heisenberg Uncertainty Principle equation, where Δx is the product of the uncertainty in position, Δp is the uncertainty in momentum of an electron, and ℏ/2 is the reduced Planck constant. (Equation created with CodeCogs online tool.)

This is a very small number that can usually be ignored, but when dealing with a particle as small as an electron, it is significant. Because of this, it becomes necessary to describe the position of an electron in terms of probability, rather than that of absolute certainty. We have to say that an electron is likely to be found within the atom in certain areas of high probability, but we cannot be 100% sure of its precise position.

Comprehension Checkpoint

The Heisenberg Uncertainty Principle tells us that

Because the electron is not a true particle, we cannot describe its movement or location in traditional terms. In other words, those that would normally be defined by simple x, y, and z coordinates. The challenges raised by the combination of wave particle duality and the Heisenberg Uncertainty Principle are simplified and expressed by Schrödinger’s equation that, when solved, produces wave functions denoted by Ψ. When Ψ is squared, the resulting solution gives the probability of finding an electron at a particular place in the atom. The Ψ2 term describes how electron density and electron probability are distributed in space around the nucleus of an atom.

The application of the Schrödinger equation is most easily understood in the case of the hydrogen atom because the one-electron atom allows us to avoid the complex interactions of multiple electrons. The time-independent form of the Schrödinger equation (Equation 2) can be solved to provide solutions that correspond to the energy levels in a hydrogen atom. In solving this equation, we find that there are multiple solutions for Ψ that we call Ψ1, Ψ2, Ψ3, etc.

Two or more atoms that bond together form a(n)
Equation 2: The time independent Schrödinger equation for hydrogen's energy levels. (Equation created with CodeCogs online tool.)

Each of these solutions has a different energy that corresponds to what we think of as different energy levels, or electron shells, within the atom. The lowest energy electron distribution is that which is closest to the nucleus, and it is called the ground state. The ground state is the most stable state of the single electron within the hydrogen atom. Other, higher energy states exist and are called excited states. As the energy of each level increases above the ground state, we refer them to them as the second, third, fourth, etc. energy levels, respectively.

Comprehension Checkpoint

The energy of the electron shells ______ as you move away from the nucleus of the atom.

Acceptable solutions to the wave equation for an electron cannot be just any values, but rather they are restricted to those that obey certain parameters. Those parameters are described by a set of three quantum numbers that are named the principal quantum number, the azimuthal quantum number, and the magnetic quantum number. These quantum numbers are given the symbols n, l, and m, respectively.

The principal quantum number n is a positive integer starting with a value of 1, where 1 corresponds to the first (lowest energy) shell known as the ground state in hydrogen. The principal quantum number then increases with the increasing energy of the shells (the excited states in hydrogen) to values of n = 2, n = 3, n = 4, etc., as one moves farther away from the nucleus to higher energy levels.

The azimuthal quantum number l (also called the orbital angular momentum quantum number) determines the physical, three-dimensional shape of the orbitals in any subshell. The value of l is dependent on the principal quantum number n, and there are multiple values of l for each value of n. Values of l are positive integers or 0, and are determined by subtracting integers from the corresponding value of n. For example, when n = 3 we can subtract 3, 2, and 1 from n to yield values for l of 0, 1, and 2, respectively. Thus, the ground state shell in hydrogen (principal quantum number 1) has only one s subshell, shell 2 has s and p subshells, and so on. The subshells indicated by l are also given the letter designations s, p, d, and f where l = 0 corresponds to an s subshell, l = 1 a p subshell, l = 2 a d subshell, and l = 3 an f subshell. Each subshell has a unique shape in 3D-space.

The magnetic quantum number m, defines the orientation (i.e., the position) and the number of orbitals within any given subshell. Values of m depend upon values of l, the azimuthal quantum number, and m can take on values of +1, -1, and all integers (including 0) in between. So for example, when l = 1, m can be +1, 0, or -1, yielding three separate orbitals, on axes x, y, and z in space. So in the case of l = 1 (the p subshell), there are three orbitals oriented in different directions in space.

Thus each energy level is given a unique set of quantum numbers to fully describe it. Only certain values for quantum numbers in any one energy level are allowed, and those combinations are summarized in Table 1. Analysis of all of the allowed solutions to the wave equation shows that orbitals can be grouped together in sets according to their l values: s, p, d, and f sets.

Table 1: The allowed solutions to the wave equation shows that orbitals can be grouped together in sets according to their l values (i.e., the s, p, d, and f sets).

n l m
(principal quantum number) azimuthal quantum number,
with letter designations)
(magnetic quantum number,
with letter designations)
1 0 (s) 0 (s)
2 0 (s),
1 (p)
0 (s),
-1/0/+1 (p)
3 0 (s), 1 (p),

2 (d)

0 (s), -1/0/+1 (p),

-2/-1/0/+1/+2 (d)

4 0 (s), 1 (p), 2 (d),

3 (f)

0 (s), -1/0/+1 (p), -2/-1/0/+1/+2 (d),

-3/-2/-1/0/+1/+2/+3 (f)

The first set of sub-orbitals are described by a type of wave function whose probability density depends only on the distance from the nucleus, and whose probability is the same in all directions. Thus they have a spherical shape and are called s orbitals. These wave functions are all found to have values of l = 0 and therefore values of m = 0, and every energy level has such a wave function starting with 1s, and moving to 2s, 3s, etc. (Figure 1).

Two or more atoms that bond together form a(n)
Figure 1: The spherical shaped s orbital. image © UC Davis ChemWiki

The second type of wave function has a probability density that depends on both the distance from the nucleus and the orientation along either the x-, y-, or z-axis in space. This leads to three, separate, equivalent (or degenerate) wave functions that have a "figure eight" shape in 3D-space. These wave functions are all found to have values of l = 1 and can take on three, different m values. These are called p orbitals and exist for every energy level except the first, thus appearing as 2p, 3p, 4p, etc. (Figure 2).

Two or more atoms that bond together form a(n)
Figure 2: The "figure eight" p orbitals. image © UC Davis ChemWiki

The third type of wave function has a probability density that depends on both the distance from the nucleus and the orientation along two of the x-, y-, and z-axes in space. This leads to five separate, equivalent wave functions. When the wave functions have equivalent energy, they are given the label "degenerate." They have complex shapes in 3D-space. These wave functions are all found to have values of l = 2 and can take on five different m values. These are called d orbitals, and every energy level except the first and second has such a wave function (Figure 3).

Two or more atoms that bond together form a(n)
Figure 3: The d orbitals, the beginning of more complex orbital shapes. image © UC Davis ChemWiki

The fourth type is a set of orbitals with even more complexity in terms of their wave function, positions, and shape in 3D-space. These wave functions are all found to have values of l = 3 and can take on seven different m values. They, too, are all degenerate. These are called f orbitals and every energy level except the first, second, and third has such a wave function, starting with 4f and moving to 5f (Figure 4).

Two or more atoms that bond together form a(n)
Figure 4: The f orbitals, which continue the complexity of shape as seen in the d orbitals. image © UC Davis ChemWiki

To summarize, the orbitals available in the first four energy levels are as follows in Table 2:

Table 2: Orbitals associated with the first four energy levels.

n orbitals
1 1s
2 2s, 2p
3 3s, 3p, 3d
4 4s, 4p, 4d, 4f

Comprehension Checkpoint

The principal quantum number n

Since orbitals are defined by very specific solutions to the Schrödinger equation, electrons must absorb very specific quantities of energy to be promoted from one energy level to another. By exposing electrons to an external source of energy, such as light, it is possible to promote the electrons from their ground state to other positions that have higher potential energies, called excited states. These ‘excited’ electrons quickly return to lower energy positions (in order to regain the stability associated with lower energies), and in doing so they release energy in specific frequencies that correspond to the energy differences between electron orbitals, or shells. Light energy is related to frequency (f) and Planck’s constant (h) in the equation E = hf.

Electrons within a hydrogen atom will absorb specific frequencies of energy that correspond to the energy gaps (ΔE) between two energy levels, thus allowing the promotion of an electron from a relatively low energy level to a relatively high energy level. Examining the absorption spectrum produced when hydrogen is irradiated, we observe a pattern of lines that supports this (Figure 5).

Two or more atoms that bond together form a(n)
Figure 5: Hydrogen's emission and absorption spectra from the Balmer series. image © Chem1 Virtual Textbook, adapted from the Online Journey through Astronomy site

Bohr proposed that electrons can transition to different positions, but only in discrete, defined steps. He thought that electrons were restricted to a specific area of space around the nucleus, much like the planets in our solar system are restricted to specific paths. The hydrogen absorption spectrum of discrete lines (rather than a continuous spectrum) shows that only very specific transitions can be made and is evidence for the quantum model.

In addition to the absorption spectrum shown above, another identical spectrum can be observed, this time with the energy being emitted rather than absorbed. In this case the electrons fall back from higher energies to lower energies rather than being promoted as in the absorption spectrum. This is called the emission spectrum of the atom and is once again made up of discrete, individual lines.

Comprehension Checkpoint

When energy is emitted, electrons

Each series of lines that appear in such spectra are named after their discoverers. The Lyman series are those lines in which the electrons are either promoted from, or to, the 1s orbital. In the Balmer, Paschen, Brackett, and Pfund series, electrons travel either from or to the orbitals where n = 2, 3, 4, and 5, respectively. The differing energy gaps (and therefore the differing frequencies) correspond to lines in different regions of the electromagnetic spectrum as a whole (Figure 6). Lyman lines are in the ultraviolet region, Balmer in the visible, and Paschen, Brackett, and Pfund in various parts (near and far) of the infrared.

Two or more atoms that bond together form a(n)
Figure 6: An energy level transition diagram for hydrogen.

In hydrogen atoms, and other single electron species such as He+ and Li2+, it is found that the energy of the orbitals is only dependent on their distance from the nucleus, with increasing energies as n increases. The lowest energy orbital is 1s, followed by 2s and 2p (that have the same energy), and 3s, 3p, and 3d (that have the same energy) etc. As such, the ground state (lowest energy state) for a hydrogen atom with only one electron is with the electron residing in the 1s orbital. The differences in energies (ΔE) required to promote electrons from one level to another can be determined by using one version of the Rydberg Formula, where RH is a constant for hydrogen and ni and nf are the initial and final energy levels respectively.

Two or more atoms that bond together form a(n)
Equation 3: The Rydberg Formula (created with CodeCogs online tool.)

Since other single electron species such as He+ and Li2+ contain more protons than the hydrogen atom, the electron experiences a greater attraction from the nucleus. As such, differing, larger amounts of energy are required for promotion of electrons in these species and a modified Rydberg constant is required in the calculation of ΔE for single electron species other than the hydrogen atom.

Comprehension Checkpoint

In atoms that have only one electron, the energy of the orbitals is dependent on the

The hydrogen atom is the simplest case to study since its single electron is both uninfluenced by other electrons and does not influence any other electrons. In this case, the wave function is relatively easy to compute. However, in more complex atoms the calculations are not so easy. For example, when two electrons are present, as in helium, things become considerably more complex. Rather than there being just one simple potential energy attraction between the nucleus and the single electron in a hydrogen atom, in helium there are now three potential energy terms to consider: the attractive forces between the nucleus and each electron, and the repulsion between the two electrons. The addition of even more electrons leads to the equation and the calculations becoming increasingly complex.

The complexity of the Schrödinger equation under these many electron circumstances means that it has to be solved by a series of approximations rather than directly. One method is to effectively apply the single electron system over and over again to produce what amounts to an approximate answer. Although not entirely correct, such approximations prove to be very workable and produce reasonable answers to the multi-electron problem.

The approximation works well enough to once again produce Ψ2 wave functions that predict the electron density in space of multi-electron atoms. The orbitals in the single electron situation and those in the multi-electron situation are still defined by the same set of restricted quantum numbers, but one difference arises regarding the specific energies of those orbitals. Previously, the energy depended only on the distance from the nucleus (that is, the energy n in multi-electron systems is found to be dependent on the value of l). This complicates the idea that relates electron energy simply to distance from the nucleus (i.e., orbital energy follows the pattern 1s < (2s = 2p) < (3s = 3p = 3d), etc.). Instead, it produces the following sequence of relative energies for the orbitals.

1s < 2s < 2p < 3s < 3p < 4s < 3d < 4p < 5s < 4d < 5p, etc.

Note the insertion of 4s and 5s earlier in the sequence than expected. This order is reflected on the periodic table. When reaching the end of period three on the periodic table (element 18, argon), we will have filled the 1s, 2s, 2p, 3s, and 3p orbitals giving a total (as expected) of 18 electrons. One would perhaps expect the next orbitals to be filled to be the 3d, since after all, they are also in the third energy level. However, the next elements on the periodic table, potassium and calcium (elements 19 and 20, respectively), fill their 4s orbitals before their 3d. Moving further along the fourth period to element 21, scandium, we find that the 3d subshell now begins to populate. When that subshell is full at element 30, zinc, we then fill the 4p subshell beginning with element 31, gallium. In short, the periodic table is ordered in such a way that it reflects the sequence of relative energies of orbitals given above.

As seen previously, the concept of electron spin introduced the need for the fourth and final quantum number, the spin quantum number, s. Its introduction ensures that each electron has a unique set of quantum numbers according to the Pauli Exclusion Principle. With there being only two values for s (+½ or -½), it follows that each orbital may only hold a maximum of two electrons.

The notation used to describe the electronic configuration of an atom involves the use of a superscript to denote the number of electrons in any given orbital. For example, in the case of beryllium that has a total of four electrons, 1s2 2s2 shows that both the 1s and the 2s orbitals are each filled with two electrons.

When identifying which orbitals the electrons in an atom occupy, we generally follow the principle referred to as the Aufbau principle, where lower energy orbitals are filled first. So, for example, 1s is filled before 2s, and 2s before 2p, etc. But what about degenerate orbitals, those that have the same energy, such as the three orbitals that exist in the 2p sub-level?

Hund’s rule states that when electrons are placed into any set of degenerate orbitals, one electron is placed into each orbital before any spin pairing takes place. This means, for example, if we consider three electrons being placed into the degenerate 2p orbitals, one would see an electronic configuration of 2px1 2py1 2pz1 (recall the p orbitals are oriented along x, y, and z axes, respectively) before we see any pairing. The fourth electron to enter the 2p sub-shell would create the need for an unavoidable pairing, hence 2px2 2py1 2pz1 (Figure 7).

Two or more atoms that bond together form a(n)
Figure 7: An illustration of Hund's Rule showing the placement of electrons in various orbitals of nitrogen (N) and oxygen (O).

Now that we have a sense of electron configuration, their energies, and their most probable positions within atoms, the next step is to describe the behavior of electrons when atoms interact with one another. Electron interaction forms the basis of a key chemical process called chemical bonding. And understanding the behavior of electrons provides us with a better understanding of the chemical behavior of the elements and their compounds.

Our Atomic Theory series continues, exploring the quantum model of the atom in greater detail. This module takes a closer look at the Schrödinger equation that defines the energies and probable positions of electrons within atoms. Using the hydrogen atom as an example, the module explains how orbitals can be described by type of wave function. Evidence for orbitals and the quantum model is provided by the absorption and emission spectra of hydrogen. Other concepts include multi-electron atoms, the Aufbau Principle, and Hund’s Rule.

Key Concepts

  • The wave-particle nature of electrons means that their position and momentum cannot be described in simple physical terms but must be described by wave functions.

  • The Schrödinger equation describes how the wave function of a wave-particle changes with time in a similar fashion to the way Newton’s second law describes the motion of a classical particle. The equation allows the calculation of each of the three quantum numbers related to individual atomic orbitals (principal, azimuthal, and magnetic).

  • The Heisenberg uncertainty principle establishes that an electron’s position and momentum cannot be precisely known together; instead we can only calculate statistical likelihood of an electron’s location.

  • The discovery of electron spin defines a fourth quantum number independent of the electron orbital but unique to an electron. The Pauli exclusion principle states that no two electrons with the same spin can occupy the same orbital.

  • Quantum numbers, when taken as a set of four (principal, azimuthal, magnetic and spin) describe acceptable solutions to the Schrödinger equation, and as such, describe the most probable positions of electrons within atoms.

  • Orbitals can be thought of as the three dimensional areas of space, defined by the quantum numbers, that describe the most probable position and energy of an electron within an atom.

  • HS-C1.4, HS-C4.4, HS-PS1.A2, HS-PS2.B3
  • Heisenberg, W. (1927). Über den anschaulichen Inhalt der quantentheoretischen Kinematik und Mechanik. Zeitschrift für Physik, 43(3–4), 172-198.

  • Pauli, W. (1925). Ueber den Einfluss der Geschwindigkeitsabhaengigkeit der Elektronenmasse auf den Zeeman-Effekt. Zeitschrift für Physik 31(1), 373-385.
  • Pauli, W. (1925). Ueber den Zusammenhang des Abschlusses der Elektronengruppen im Atom mit der Komplexstruktur der Spektren. Zeitschrift für Physik 31(1), 765-783.
  • Pauli, W. (1946). Remarks on the history of the Exclusion Principle. Science, New Series, 103(2669), 213-215.
  • Schrödinger, E. (1926). Quantisierung als Eigenwertproblem. Annalen der Physik, 384(4), 361-376.
  • Stoner, E.C. (1924). The distribution of electrons among atomic energy levels. The London, Edinburgh and Dublin Philosophical Magazine (6th series), 48, 719-736.

Adrian Dingle, B.Sc., Anthony Carpi, Ph.D. “Atomic Theory IV” Visionlearning Vol. CHE-3 (7), 2016.

Top


Page 13

Physical States and Properties

by Megan Cartwright, Ph.D.

If you’ve ever changed an old incandescent light bulb, you might have noticed what looks like black powder coating the inside of the bulb. That black coating is actually metal atoms that escaped from the bulb’s tungsten filament and condensed on its glass (Figure 1). While this little bit of tungsten residue is annoying for modern people who like to read at night, in the early 1900s light bulbs used to burn out their filaments and turn black very quickly. Then in 1913, the American chemist Irving Langmuir figured out a surprising solution to keep bulbs burning bright: fill the bulb with an inert, non-toxic gas called argon.

Two or more atoms that bond together form a(n)
Figure 1: A new, clear light bulb compared with a blackened one. The black coating is from metal atoms that escaped from the bulb’s tungsten filament and condensed on its glass. image © new bulb, Dave Gough / burned-out bulb, Paul Cowan

Before Langmuir, manufacturers made light bulbs with a vacuum inside to prevent oxygen from contacting the filament. This was because when current ran through the filament, it heated to 3,000°C—hot enough to oxidize the metal in the filament. While this temperature helpfully caused the filament to radiate visible light, it occasionally caused a tungsten atom to sublime (change directly from solid to gas phase) off the filament and onto the bulb’s glass, deteriorating the filament and blackening the bulb.

Langmuir figured out that by filling the bulb with argon gas, the tungsten atoms would take much longer to blacken the bulb. Instead of streaking straight towards the glass walls, they would collide and bounce off the argon atoms, sometimes even ricocheting back into the filament.

Langmuir was able to solve the problem of blackening light bulbs because he was familiar with kinetic-molecular theory (KMT). By making several assumptions about the motion and energy of molecules, KMT provides scientists with a useful framework for understanding how the behavior of molecules influences the behaviors of different states of matter, particularly the gas state. As the story of Langmuir’s light bulbs shows, this framework can be a useful tool for understanding and solving real-world problems. But KMT hasn’t always existed: When Langmuir figured out how to make light bulbs last longer in 1913, he was relying on many centuries of work by scientists who had developed the assumptions at the core of modern KMT.

In the 17th century, the Italian mathematician Evangelista Torricelli built the first mercury barometer by filling a glass tube sealed at one end with mercury and then inverting the open end into a tub full of the liquid metal. To the surprise of his contemporaries, the tube remained partially filled—almost as if something was pushing down on the mercury in the tub, and forcing the liquid metal up the tube (Figure 2). Most significantly, the level at which mercury rose in the tube changed from day to day, challenging scientists to explain how mercury was forced up the closed glass tube.

Two or more atoms that bond together form a(n)
Figure 2: An example of Evangelista Torricelli's experiment with a mercury barometer, where he filled a glass tube sealed at one end with mercury and then inverted the open end into a tub full of the liquid metal. The tube of mercury remained partially filled even though inverted. image © Technica Curiosa [...] by Gasparis Schott

The British scientists Robert Boyle and Robert Hooke devised an experiment to figure out what was pressing down on the mercury. Working with a candy cane-shaped glass tube that had its short leg sealed off, Boyle poured in just enough mercury to fill the tube’s curve and trap air inside the short leg. When he poured in still more mercury, Boyle saw that the while the volume of trapped air shrank, the air somehow pushed against the mercury and forced it back up the long leg. Boyle reasoned that air must also weigh down on the mercury in the Torricelli’s tub and exert the force driving mercury partially up the tube (to learn more about Boyle’s experiment, see our module Properties of Gases).

But how could air—which many scientists had regarded as an indivisible element—weigh down on mercury? By the early 18th century, scientists realized that air was made up of tiny particles. However, these same scientists couldn’t imagine air particles just floating in space. They assumed that the particles vibrated and spun while being held in place by an invisible substance called ether.

The Swiss mathematician Daniel Bernoulli had a different idea about how particles could be suspended in air. In his 1738 book Hydrodynamica, Bernoulli sketched out a thought experiment illustrating how the linear motion of air particles could exert pressure. Bernoulli first asked his readers to imagine a cylinder fitted with a movable piston. Next, he directed the reader to picture the air particles as tiny spheres that zipped around in all directions, colliding with each other and the piston. These numerous, constant collisions would “kick” the piston aloft. Furthermore, Bernoulli suggested that if the air was heated, the particles would zoom faster, striking the piston more often and kicking the piston still higher in the cylinder.

With his thought experiment, Bernoulli was the first to develop several assumptions about molecules and heat that are integral to modern KMT. Like modern KMT, Bernoulli assumed that molecules behave like tiny spheres in constant linear motion. Working from this assumption, he reasoned that the molecules would constantly collide with each other and the walls of a container, thereby exerting pressure on these walls. Importantly, he also assumed that heat affects the movement of molecules.

Though largely correct, Bernoulli’s ideas were mostly dismissed by contemporary scientists. He didn’t offer any experimental data to support his ideas. Furthermore, accepting his ideas required a literal belief in atoms, which many scientists were skeptical existed up to the 19th century. Equating motion with temperature implied that there must be an absolute minimum temperature at which point all motion ceased.

And finally, Bernoulli’s kinetic theory competed with caloric theory, a prominent idea at the time. According to caloric theory, caloric was a “heat substance” that engulfed gas molecules, causing them to repel each other so forcefully they would shoot into the walls of a container. This competing idea about what caused air pressure was championed by influential chemists such as Antoine Lavoisier and John Dalton, while Bernoulli’s idea was largely ignored until the 19th century. Unfortunately, it is not uncommon for good ideas to take time to be accepted in science as previously accepted theories are disproven. However, over time, scientific progress assures that theories which better explain the data collected take hold, as Bernoulli’s did.

Comprehension Checkpoint

Like modern kinetic-molecular theory, Bernoulli theorized that

Unlike Lavoisier and Dalton, the 19th century German physicist Rudolf Clausius rejected caloric theory. Instead of regarding heat as a substance that surrounds molecules, Clausius proposed that heat is a form of energy that affects the temperature of matter by changing the motion of molecules in matter. This kinetic theory of heat enabled Clausius to study and predict the flow of heat—a field we now call thermodynamics (for more information, see our module Thermodynamics I).

In his 1857 paper, “On the nature of the motion which we call heat,” Clausius speculated on how heat energy, temperature, and molecule motion could explain gas behavior. In doing this, he proposed several ideas about the molecules of gases. These ideas have come to be accepted for ideal gases—theoretical gases that perfectly obey the ideal gas equation (for more information, see our Properties of Gases module). Clausius proposed that the space taken up by ideal gas molecules should be regarded as infinitesimal when compared to the space occupied by the whole gas – in other words, a gas consists mostly of empty space. Second, he suggested that the intermolecular forces between molecules should be treated as infinitesimal.

A key part of Clausius’s ideas was his work on the mathematical relationship between heat, temperature, molecule motion, and kinetic energy—the energy of motion. He proposed that the net kinetic energy of the molecules in an ideal gas is directly proportional to the gas’s absolute temperature, T. Its kinetic energy, Ek, is therefore determined by the number of gas molecules, n, which each have a molecule mass of m, and are moving with velocity u, as shown below:

With this equation and observational data on the weights and volumes of gases at specific temperatures, Clausius was able to calculate the average speeds of gas molecules such as oxygen (an astonishing 461 m/s!). However, the Dutch meteorologist Christoph Hendrik Diederik Buys-Ballot quickly pointed out a problem with these speed calculations. If gas molecules moved hundreds of meters per second, then an odorous gas (like in perfume) should spread across a room almost instantly. Instead, perfumes and other scents usually took several minutes to reach people across a room. This suggested that either Clausius’s mathematical relationship was wrong, or that something more complicated was happening with real gas molecules.

Comprehension Checkpoint

Clausius proposed that

Buys-Ballot’s objection forced Clausius to re-think his ideas about gas molecules. If a gas molecule could travel at 461 m/s, but still take minutes to cross a room, it must be encountering lots of obstacles—such as other gas molecules. Clausius realized then that one of his core ideas about ideal gas molecules had to change.

In 1859, Clausius published a paper proposing that, instead of being infinitesimally small, gas molecules had to be big enough that it could collide with another molecule, and there had to be so many fast-moving gas molecules present that it couldn’t travel far before doing so. The average distance that a molecule travels between collisions has come to be known as its mean free path. Clausius realized that while the mean free path must be very big compared to the actual size of the molecules, it would still have to be small enough that a fast-moving molecule would collide with other molecules many times each second (Figure 3).

Two or more atoms that bond together form a(n)
Figure 3: A simplified illustration of a molecule's (blue dot) mean free path, which is the average distance that a molecule travels between collisions (dotted lines). The solid line indicates the actual distance between the beginning and end of the molecule's journey.

Thus, gas molecules are constantly colliding and changing directions. While it’s tempting to picture molecules colliding every few seconds like bumper cars at an amusement park, the collisions are so much more frequent. For example, at room temperature, one oxygen molecule travels an average distance of 67 nm (almost 1,500 times narrower than the width of a human hair) before colliding with another molecule. And this single molecule collides with others 7.2 billion times per second! This astounding frequency of collisions explains how gas molecules can zip along at hundreds of meters per second, but still take minutes to cross a room.

Clausius’s idea about mean free path was vital to how Langmuir solved the blackening light bulb problem. Thanks to Clausius, Langmuir understood that he needed to decrease the mean free path for the tungsten atoms sublimating off the filament. In a vacuum, the mean free path was very long, and the tungsten atoms quickly make their way from the filament to the inside of the bulb. By adding an inert gas like argon, Langmuir increased the number of molecules in the bulb and the frequency of collisions—thereby decreasing the mean free path, and increasing the bulb’s life.

Clausius’s ideas about mean free path, molecule motion, and kinetic energy intrigued James Clerk Maxwell, a contemporary Scottish physicist. In 1860, Maxwell published his first kinetic theory paper expanding on Clausius’s work. Building on Clausius’s calculations for the average speed of oxygen and other molecules, Maxwell further developed the idea that the gas molecules in a sample of gas are moving at different speeds. Within this range of possible speeds, some molecules are slower than the average and some faster (Figure 4); furthermore, a molecule’s speed can change when it collides with another molecule. Only a tiny number of the gas molecules are actually moving at the slowest and fastest speeds possible—but we know now that this small number of speedy molecules are especially important, because they are the most likely molecules to undergo a chemical reaction.

Two or more atoms that bond together form a(n)
Figure 4: Since heavier molecules move more slowly than lighter molecules, the heavier molecules (xenon, argon) have a narrow speed distribution and the lighter molecules (neon, helium) have a spread out speed distribution. image © Pdbailey

Along with these ideas, Maxwell proposed that gas particles should be treated mathematically as spheres that undergo perfectly elastic collisions. This means that the net kinetic energy of the spheres is the same before and after they collide, even if their velocities change. Together, Clausius and Maxwell developed several key assumptions that are vital to modern KMT.

Comprehension Checkpoint

The average distance that a molecule travels between collisions is known as

A major use of modern KMT is as a framework for understanding gases and predicting their behavior. KMT links the microscopic behaviors of ideal gas molecules to the macroscopic properties of gases. In its current form, KMT makes five assumptions about ideal gas molecules:

  1. Gases consist of many molecules in constant, random, linear motion.
  2. The volume of all the molecules is negligible compared to the gas’s total volume.
  3. Intermolecular forces are negligible.
  4. The average kinetic energy of all molecules does not change, so long as the gas’s temperature is constant. In other words, collisions between molecules are perfectly elastic.
  5. The average kinetic energy of all molecules is proportional to the absolute temperature of the gas. This means that, at any temperature, gas molecules in equilibrium have the same average kinetic energy (but NOT the same velocity and mass).

With KMT’s assumptions, scientists are able to describe on a molecular level the behaviors of gases. These behaviors are common to all gases because of the relationships between gas pressure, volume, temperature, and amount, which are described and predicted by the gas laws (for more on the gas laws, please see our Properties of Gases module). But KMT and the gas laws are useful for understanding more than abstract ideas about chemistry. With KMT and the gas laws, we can better understand the behaviors of real gases, such as the air we use to inflate tires, as we’ll explore more below.

Comprehension Checkpoint

Gases consist of many molecules in _____, random, linear motion.

Boyle’s law describes how, for a fixed amount of gas, its volume is inversely proportional to its pressure (Figure 5). This means that if you took all the air from a fully inflated bike tire and put the air inside a much larger, empty car tire, the air would not be able to exert enough pressure to inflate the car tire. While this example about the relationship between gas volume and pressure may seem intuitive, KMT can help us understand the relationship on a molecular level. According to KMT, air pressure depends on how often and how forcefully air molecules collide with tire walls. So when the volume of the container increases (like when we transfer air from the bike tire to the car tire), the air molecules have to travel farther before they can collide with the tire’s walls. This means that there are fewer collisions per unit of time, which results in lower pressure (and an underinflated car tire).

Two or more atoms that bond together form a(n)
Figure 5: Boyle's law states that so long as temperature is kept constant, the volume of a fixed amount of gas is inversely proportional to the pressure placed on the gas.

Charles’s law describes how a gas’s volume is directly proportional to its absolute temperature (Figure 6)—and also why your car tire pressure increases the longer you drive (and thus why you should always measure tire pressure after your car has been parked for a long period). Using KMT, we can understand that as the friction between the tires and road raises the air temperature inside the tires, the air molecules’ kinetic energy and speed are also increasing. Because the molecules are zipping around faster, they collide more frequently and more forcefully with the tire’s walls, thereby increasing the pressure. Both Charles’s law and KMT also explain why tire pressure decreases after you park. As the tires cool down, the air molecules move more slowly and collide less with the tire’s walls, thereby exerting less pressure.

Two or more atoms that bond together form a(n)
Figure 6: Charles's law states that when pressure is kept constant, a fixed amount of gas linearly increases its volume as its temperature increases.

While KMT is a useful tool for understanding the linked behaviors of molecules and matter, particularly gases, KMT does have limitations related to how its theoretical assumptions differ from the behavior of real matter. In particular, KMT’s assumptions that intermolecular forces are negligible, and the volume of molecules is negligible, aren’t always valid. Real gas molecules do experience intermolecular forces. As pressure on a real gas increases and forces its molecules closer together, the molecules can attract one another. This attraction slows down the molecules just a little bit before they slam into one another or the walls of a container, so that the pressure inside a container of real gas molecules is slightly lower than we would expect based on KMT. These intermolecular forces are particularly influential when gas molecules are moving more slowly, such as at low temperature.

While growing pressure on a real gas initially allows its intermolecular forces to have more influence, a different factor gains more influence as the pressure continues to grow. While KMT assumes that gas molecules have no volume, real gas molecules do have volume. This gives a real gas greater volume at high pressure than would be predicted from KMT. Furthermore, as a real gas is compressed, the mean free path of its molecules decreases and the molecules collide more often—thereby increasing the pressure exerted by a real gas compared to KMT’s prediction.

Ultimately, KMT is most useful and accurate when gases are under conditions that cause molecules to behave consistently with KMT’s assumptions. These conditions often happen at low pressure, where molecules have lots of empty space to move in, and the molecule volumes are very small compared to the total volume. And the conditions often occur at high temperature, when the molecules possess a high kinetic energy and fast speed, which lets them overcome the attractive forces between molecules.

Ultimately, KMT provides assumptions about molecule behavior that can be used both as the basis for other theories about molecules, and to solve real-world problems. Clausius’ concept of a molecule’s mean free path underlies our modern ideas of diffusion and Brownian motion, which can explain why scents from perfume or baking cookies take so long to cross a room. And while Langmuir used KMT’s assumptions to develop a longer-lasting light bulb, still other scientists are deploying their knowledge of KMT in more dramatic and controversial ways. By understanding how real gas molecules behave and move, scientists are able to separate gas molecules from each other based on tiny differences in mass—a key principle behind, for example, how uranium isotopes are enriched for use in nuclear weapons.

Over four hundred years, scientists including Rudolf Clausius and James Clerk Maxwell developed the kinetic-molecular theory (KMT) of gases, which describes how molecule properties relate to the macroscopic behaviors of an ideal gas—a theoretical gas that always obeys the ideal gas equation. KMT provides assumptions about molecule behavior that can be used both as the basis for other theories about molecules and to solve real-world problems.

Key Concepts

  • Kinetic-molecular theory states that molecules have an energy of motion (kinetic energy) that depends on temperature.

  • Rudolf Clausius developed the kinetic theory of heat, which relates energy in the form of heat to the kinetic energy of molecules.

  • Over four hundred years, scientists have developed the kinetic-molecular theory of gases, which describes how molecule properties relate to the macroscopic behaviors of an ideal gas—a theoretical gas that always obeys the ideal gas equation.

  • The kinetic-molecular theory of gases assumes that ideal gas molecules (1) are constantly moving; (2) have negligible volume; (3) have negligible intermolecular forces; (4) undergo perfectly elastic collisions; and (5) have an average kinetic energy proportional to the ideal gas’s absolute temperature.

  • Bruce Fye, W. (2001). Johann and Daniel Bernoulli. Clinical Cardiology, 24(9): 634-635.

  • Brush, S.G. (1999). Gadflies and geniuses in the history of gas theory. Synthese, 119(1): 11-43.
  • Clausius, R. (1857). Ueber die Art der Bewegung, welche wir Wärme nennen. Annalen der Physik, 176(3): 353-380.
  • Cornely-Moss, K. (1995). Kinetic theory of gases. Journal of Chemical Education, 72(8): 715.
  • Garber, E. (1871). Subjects great and small: Maxwell on Saturn's rings and kinetic theory. Philosophical Transactions of the Royal Society of London A: Mathematical, Physical and Engineering Sciences, 366(1871): 697-1705.
  • Jennings, S.G. (1988). The mean free path in air. Journal of Aerosol Science, 19(2): 159-166.

Megan Cartwright, Ph.D. “Kinetic-Molecular Theory” Visionlearning Vol. CHE-4 (2), 2017.

Top


Page 14

Reactions and Changes

by Anthony Carpi, Ph.D.

Traditional chemical reactions occur as a result of the interaction between valence electrons around an atom's nucleus (see our Chemical Reactions module for more information). In 1896, Henri Becquerel expanded the field of chemistry to include nuclear changes when he discovered that uranium emitted radiation. Soon after Becquerel's discovery, Marie Sklodowska Curie began studying radioactivity and completed much of the pioneering work on nuclear changes. Curie found that radiation was proportional to the amount of radioactive element present, and she proposed that radiation was a property of atoms (as opposed to a chemical property of a compound). Marie Curie was the first woman to win a Nobel Prize and the first person to win two (the first, shared with her husband Pierre Curie and Henri Becquerel for discovering radioactivity; the second for discovering the radioactive elements radium and polonium).

In 1902, Frederick Soddy proposed the theory that "radioactivity is the result of a natural change of an isotope of one element into an isotope of a different element." Nuclear reactions involve changes in particles in an atom's nucleus and thus cause a change in the atom itself. All elements heavier than bismuth (Bi) (and some lighter) exhibit natural radioactivity and thus can "decay" into lighter elements. Unlike normal chemical reactions that form molecules, nuclear reactions result in the transmutation of one element into a different isotope or a different element altogether (remember that the number of protons in an atom defines the element, so a change in protons results in a change in the atom). There are three common types of radiation and nuclear changes:

Alpha Radiation (α) is the emission of an alpha particle from an atom's nucleus. An α particle contains two protons and two neutrons (and is similar to a He nucleus:

Two or more atoms that bond together form a(n)
). When an atom emits an a particle, the atom's atomic mass will decrease by four units (because two protons and two neutrons are lost) and the atomic number (z) will decrease by two units. The element is said to "transmutate" into another element that is two z units smaller. An example of an α transmutation takes place when uranium decays into the element thorium (Th) by emitting an alpha particle, as depicted in the following equation:

Two or more atoms that bond together form a(n)

(Note: in nuclear chemistry, element symbols are traditionally preceded by their atomic weight [upper left] and atomic number [lower left].)

Beta Radiation (β) is the transmutation of a neutron into a proton and an electron (followed by the emission of the electron from the atom's nucleus:

Two or more atoms that bond together form a(n)
). When an atom emits a β particle, the atom's mass will not change (since there is no change in the total number of nuclear particles); however, the atomic number will increase by one (because the neutron transmutated into an additional proton). An example of this is the decay of the isotope of carbon called carbon-14 into the element nitrogen:

Two or more atoms that bond together form a(n)

Gamma Radiation (γ) involves the emission of electromagnetic energy (similar to light energy) from an atom's nucleus. No particles are emitted during gamma radiation, and thus gamma radiation does not itself cause the transmutation of atoms; however, γ radiation is often emitted during, and simultaneous to, α or β radioactive decay. X-rays, emitted during the beta decay of cobalt-60, are a common example of gamma radiation.

Comprehension Checkpoint

Radiation can result in an atom having a different atomic number.

Radioactive decay proceeds according to a principle called the half-life. The half-life (T½) is the amount of time necessary for one-half of the radioactive material to decay. For example, the radioactive element bismuth (210Bi) can undergo alpha decay to form the element thallium (206Tl) with a reaction half-life equal to five days. If we begin an experiment starting with 100 g of bismuth in a sealed lead container, after five days we will have 50 g of bismuth and 50 g of thallium in the jar. After another five days (ten from the starting point), one-half of the remaining bismuth will decay and we will be left with 25 g of bismuth and 75 g of thallium in the jar. As illustrated, the reaction proceeds in halves, with half of whatever is left of the radioactive element decaying every half-life period.

Two or more atoms that bond together form a(n)
Radioactive Decay of Bismuth-210 (T½ = 5 days)

The fraction of parent material that remains after radioactive decay can be calculated using the equation:

Fraction remaining =   1 
2n
(where n = # half-lives elapsed)

The amount of a radioactive material that remains after a given number of half-lives is therefore:

Amount remaining = Original amount * Fraction remaining


The decay reaction and T½ of a substance are specific to the isotope of the element undergoing radioactive decay. For example, Bi210 can undergo a decay to Tl206 with a T½of five days. Bi215, by comparison, undergoes bdecay to Po215 with a T½ of 7.6 minutes, and Bi208 undergoes yet another mode of radioactive decay (called electron capture) with a T½ of 368,000 years!

Comprehension Checkpoint

All radioactive material decays at the same rate.

While many elements undergo radioactive decay naturally, nuclear reactions can also be stimulated artificially. Although these reactions also occur naturally, we are most familiar with them as stimulated reactions. There are two such types of nuclear reactions:

1) Nuclear fission: reactions in which an atom's nucleus splits into smaller parts, releasing a large amount of energy in the process. Most commonly this is done by "firing" a neutron at the nucleus of an atom. The energy of the neutron "bullet" causes the target element to split into two (or more) elements that are lighter than the parent atom.

Two or more atoms that bond together form a(n)
The Fission Reaction of Uranium-235

Two or more atoms that bond together form a(n)

Interactive Animation: The Fission of U235

During the fission of U235, three neutrons are released in addition to the two daughter products. If these released neutrons collide with nearby U235 nuclei, they can stimulate the fission of these atoms and start a self-sustaining nuclear chain reaction. This chain reaction is the basis of nuclear power. As uranium atoms continue to split, a significant amount of energy is released from the reaction. The heat released during this reaction is harvested and used to generate electrical energy.

Two or more atoms that bond together form a(n)

Interactive Animation: Two Types of Nuclear Chain Reactions

Nuclear fusion: reactions in which two or more elements "fuse" together to form one larger element, releasing energy in the process. A good example is the fusion of two "heavy" isotopes of hydrogen (deuterium: H2 and tritium: H3) into the element helium.

Two or more atoms that bond together form a(n)
Nuclear Fusion of Two Hydrogen Isotopes

Two or more atoms that bond together form a(n)

Interactive Animation: Nuclear Fusion

Fusion reactions release tremendous amounts of energy and are commonly referred to as thermonuclear reactions. Although many people think of the sun as a large fireball, the sun (and all stars) are actually enormous fusion reactors. Stars are primarily gigantic balls of hydrogen gas under tremendous pressure due to gravitational forces. Hydrogen molecules are fused into helium and heavier elements inside of stars, releasing energy that we receive as light and heat.

Beginning with the work of Marie Curie and others, this module traces the development of nuclear chemistry. It describes different types of radiation: alpha, beta, and gamma. The module then applies the principle of half-life to radioactive decay and explains the difference between nuclear fission and nuclear fusion.

  • HS-C5.5, HS-PS1.C1, HS-PS3.A1

Anthony Carpi, Ph.D. “Nuclear Chemistry” Visionlearning Vol. CHE-2 (3), 2003.

Top


Page 15

To understand life as we know it, we must first understand a little bit of organic chemistry. Organic molecules contain both carbon and hydrogen. Though many organic chemicals also contain other elements, it is the carbon-hydrogen bond that defines them as organic. Organic chemistry defines life. Just as there are millions of different types of living organisms on this planet, there are millions of different organic molecules, each with different chemical and physical properties. There are organic chemicals that make up your hair, your skin, your fingernails, and so on. The diversity of organic chemicals is due to the versatility of the carbon atom. Why is carbon such a special element? Let's look at its chemistry in a little more detail.

Carbon (C) appears in the second row of the periodic table and has four bonding electrons in its valence shell (see our Periodic Table module for more information). Similar to other non-metals, carbon needs eight electrons to satisfy its valence shell. Carbon therefore forms four bonds with other atoms (each bond consisting of one of carbon's electrons and one of the bonding atom's electrons). Every valence electron participates in bonding; thus, a carbon atom's bonds will be distributed evenly over the atom's surface. These bonds form a tetrahedron (a pyramid with a spike at the top), as illustrated below:

Two or more atoms that bond together form a(n)
Carbon forms 4 bonds

Organic chemicals get their diversity from the many different ways carbon can bond to other atoms. The simplest organic chemicals, called hydrocarbons, contain only carbon and hydrogen atoms; the simplest hydrocarbon (called methane) contains a single carbon atom bonded to four hydrogen atoms:

Two or more atoms that bond together form a(n)
Methane - a carbon atom bonded to 4 hydrogen atoms 

But carbon can bond to other carbon atoms in addition to hydrogen, as illustrated in the molecule ethane below:

Two or more atoms that bond together form a(n)
Ethane - a carbon-carbon bond

In fact, the uniqueness of carbon comes from the fact that it can bond to itself in many different ways. Carbon atoms can form long chains:

Two or more atoms that bond together form a(n)
Hexane - a 6-carbon chain

branched chains:

Two or more atoms that bond together form a(n)
Isohexane - a branched-carbon chain

rings:

Two or more atoms that bond together form a(n)
Cyclohexane - a ringed hydrocarbon

There appears to be almost no limit to the number of different structures that carbon can form.  To add to the complexity of organic chemistry, neighboring carbon atoms can form double and triple bonds in addition to single carbon-carbon bonds:

Two or more atoms that bond together form a(n)
Two or more atoms that bond together form a(n)
Two or more atoms that bond together form a(n)
Single bonding Double bonding Triple bonding

Keep in mind that each carbon atom forms four bonds. As the number of bonds between any two carbon atoms increases, the number of hydrogen atoms in the molecule decreases (as can be seen in the figures above).

Comprehension Checkpoint

__________ can form long chains, branched chains, and rings.

The simplest hydrocarbons are those that contain only carbon and hydrogen. These simple hydrocarbons come in three varieties depending on the type of carbon-carbon bonds that occur in the molecule.

Alkanes are the first class of simple hydrocarbons and contain only carbon-carbon single bonds. The alkanes are named by combining a prefix that describes the number of carbon atoms in the molecule with the root ending "ane". The names and prefixes for the first ten alkanes are given in the following table.

Carbon
Atoms
Prefix Alkane
Name
Chemical
Formula
Structural
Formula
1 Meth Methane CH4 CH4
2 Eth Ethane C2H6 CH3CH3
3 Prop Propane C3H8 CH3CH2CH3
4 But Butane C4H10 CH3CH2CH2CH3
5 Pent Pentane C5H12 CH3CH2CH2CH2CH3
6 Hex Hexane C6H14 ...
7 Hept Heptane C7H16
8 Oct Octane C8H18
9 Non Nonane C9H20
10 Dec Decane C10H22

The chemical formula for any alkane is given by the expression CnH2n+2. The structural formula, shown for the first five alkanes in the table, shows each carbon atom and the elements that are attached to it. This structural formula is important when we begin to discuss more complex hydrocarbons. The simple alkanes share many properties in common. All enter into combustion reactions with oxygen to produce carbon dioxide and water vapor. In other words, many alkanes are flammable. This makes them good fuels. For example, methane is the main component of natural gas, and butane is common lighter fluid.

CH4 + 2O2 CO2 + 2H20
The chemical reaction between a fuel (for example wood) and an oxidizing agent.

The second class of simple hydrocarbons, the alkenes, consists of molecules that contain at least one double-bonded carbon pair. Alkenes follow the same naming convention used for alkanes. A prefix (to describe the number of carbon atoms) is combined with the ending "ene" to denote an alkene. Ethene, for example is the two-carbon molecule that contains one double bond. The chemical formula for the simple alkenes follows the expression CnH2n. Because one of the carbon pairs is double bonded, simple alkenes have two fewer hydrogen atoms than alkanes.

Two or more atoms that bond together form a(n)
Ethene

Alkynes are the third class of simple hydrocarbons and are molecules that contain at least one triple-bonded carbon pair. Like the alkanes and alkenes, alkynes are named by combining a prefix with the ending "yne" to denote the triple bond. The chemical formula for the simple alkynes follows the expression CnH2n-2.

Two or more atoms that bond together form a(n)
Ethyne

Comprehension Checkpoint

The simplest of hydrocarbons are called

Because carbon can bond in so many different ways, a single molecule can have different bonding configurations. Consider the two molecules illustrated here:

C6H14
Two or more atoms that bond together form a(n)

CH3CH2CH2CH2CH2CH3

C6H14
Two or more atoms that bond together form a(n)
CH3
I
CH3 CH2 CH CH2 CH3

Both molecules have identical chemical formulas (shown in the left column); however, their structural formulas (and thus some chemical properties) are different. These two molecules are called isomers. Isomers are molecules that have the same chemical formula but different structural formulas.

Comprehension Checkpoint

When molecules have the same number and type of atoms, they must have the same structure.

In addition to carbon and hydrogen, hydrocarbons can also contain other elements. In fact, many common groups of atoms can occur within organic molecules, these groups of atoms are called functional groups. One good example is the hydroxyl functional group. The hydroxyl group consists of a single oxygen atom bound to a single hydrogen atom (-OH). The group of hydrocarbons that contain a hydroxyl functional group is called alcohols. The alcohols are named in a similar fashion to the simple hydrocarbons, a prefix is attached to a root ending (in this case "anol") that designates the alcohol. The existence of the functional group completely changes the chemical properties of the molecule. Ethane, the two-carbon alkane, is a gas at room temperature; ethanol, the two-carbon alcohol, is a liquid.

Two or more atoms that bond together form a(n)
Ethanol

Ethanol, common drinking alcohol, is the active ingredient in "alcoholic" beverages such as beer and wine.

The chemical basis of all living organisms is linked to the way that carbon bonds with other atoms. This introduction to organic chemistry explains the many ways that carbon and hydrogen form bonds. Basic hydrocarbon nomenclature is described, including alkanes, alkenes, alkynes, and isomers. Functional groups of atoms within organic molecules are discussed.

Anthony Carpi, Ph.D. “Carbon Chemistry” Visionlearning Vol. CHE-2 (4), 2003.


Page 16

Reactions and Changes

by Anthony Carpi, Ph.D.

Top