Physical Science Revised
Physical Science Revised
Physical Science
Physical Science
Module 1
Week 1 and 2
Most Essential Learning Competencies
1. Give evidence for and describe the formation of heavier elements during star
formation and evolution.
2. Explain how the concept of atomic number led to the synthesis of new elements
in the laboratory.
Lecture
Throughout the Milky Way Galaxy (and even near the Sun itself), astronomers have
discovered stars that are well evolved or even approaching extinction, or both, as well
as occasional stars that must be very young or still in the process of formation.
PHYSI
Evolutionary effects on these stars are not negligible, even for a middle-aged star such
as the Sun. More massive stars must display more spectacular effects because the rate
of conversion of mass into energy is higher. While the Sun produces energy at the rate
CAL
of about two ergs per gram per second, a more luminous main-sequence star can
release energy at a rate some 1,000 times greater. Consequently, effects that require
billions of years to be easily recognized in the Sun might occur within a few million years
SCIEN
in highly luminous and massive stars. A supergiant star such as Antares, a bright main-
sequence star such as Rigel, or even a more modest star such as Sirius cannot have
endured as long as the Sun has endured. These stars must have been formed relatively
CE
recently.
Stellar evolution
Birth of stars and evolution to the main sequence
Detailed radio maps of nearby molecular clouds reveal that they are clumpy, with
regions containing a wide range of densities—from a few tens
of molecules (mostly hydrogen) per cubic centimeter to more than one million. Stars
form only from the densest regions, termed cloud cores, though they need not lie at the
geometric center of the cloud. Large cores (which probably contain subcondensations)
up to a few light-years in size seem to give rise to unbound associations of very
massive stars (called OB associations after the spectral type of their most prominent
members, O and B stars) or to bound clusters of less massive stars. Whether a stellar
Physical Science
Center of the Orion Nebula (M42) Astronomers have identified some 700 young stars in
this 2.5-light-year-wide area. They have also detected over 150 protoplanetary disks, or
proplyds, which are believed to be embryonic solar systems that will eventually form
planets. These stars and proplyds generate most of the nebula's light. This picture is a
mosaic combining 45 images taken by the Hubble Space Telescope.
Low-mass stars also are formed in associations called T associations after the
prototypical stars found in such groups, T Tauri stars. The stars of a T association form
from loose aggregates of small molecular cloud cores a few tenths of a light-year in size
that are randomly distributed through a larger region of lower average density. The
formation of stars in associations is the most common outcome; bound clusters account
for only about 1 to 10 percent of all star births. The overall efficiency of star formation in
associations is quite small. Typically less than 1 percent of the mass of a molecular
cloud becomes stars in one crossing time of the molecular cloud (about 5 106 years).
Low efficiency of star formation presumably explains why any interstellar gas remains in
the Galaxy after 1010 years of evolution. Star formation at the present time must be a
mere trickle of the torrent that occurred when the Galaxy was young.
Physical Science
The W5 Star Formation Region in an image taken by the Spitzer Space Telescope.
A typical cloud core rotates fairly slowly, and its distribution of mass is strongly
concentrated toward the center. The slow rotation rate is probably attributable to the
braking action of magnetic fields that thread through the core and its envelope. This
magnetic braking forces the core to rotate at nearly the same angular speed as the
envelope as long as the core does not go into dynamic collapse. Such braking is an
important process because it assures a source of matter of relatively low angular
momentum (by the standards of the interstellar medium) for the formation of stars and
planetary systems. It also has been proposed that magnetic fields play an important role
in the very separation of the cores from their envelopes. The proposal involves the
slippage of the neutral component of a lightly ionized gas under the action of the self-
gravity of the matter past the charged particles suspended in a background magnetic
field. This slow slippage would provide the theoretical explanation for the observed low
overall efficiency of star formation in molecular clouds.
At some point in the course of the evolution of a molecular cloud, one or more of its
cores become unstable and subject to gravitational collapse. Good arguments exist that
the central regions should collapse first, producing a condensed protostar whose
contraction is halted by the large buildup of thermal pressure when radiation can no
longer escape from the interior to keep the (now opaque) body relatively cool. The
protostar, which initially has a mass not much larger than Jupiter, continues to grow by
accretion as more and more overlying material falls on top of it. The infall shock, at the
surfaces of the protostar and the swirling nebular disk surrounding it, arrests the inflow,
creating an intense radiation field that tries to work its way out of the in falling envelope
of gas and dust. The photons, having optical wavelengths, are degraded into longer
wavelengths by dust absorption and reemission, so that the protostar is apparent to a
distant observer only as an infrared object. Provided that proper account is taken of the
effects of rotation and magnetic field, this theoretical picture correlates with the radiative
spectra emitted by many candidate protostars discovered near the centers of molecular
cloud cores.
An interesting speculation concerning the mechanism that ends the in fall phase exists:
it notes that the inflow process cannot run to completion. Since molecular clouds as a
whole contain much more mass than what goes into each generation of stars, the
depletion of the available raw material is not what stops the accretion flow. A rather
different picture is revealed by observations at radio, optical, and X-ray wavelengths. All
newly born stars are highly active, blowing powerful winds that clear the surrounding
regions of the in falling gas and dust. It is apparently this wind that reverses the
accretion flow.
Physical Science
The geometric form taken by the outflow is intriguing. Jets of matter seem to squirt in
opposite directions along the rotational poles of the star (or disk) and sweep up the
ambient matter in two lobes of outwardly moving molecular gas—the so-called bipolar
outflows. Such jets and bipolar outflows are doubly interesting because their
counterparts were discovered sometime earlier on a fantastically larger scale in the
double-lobed forms of extragalactic radio sources, such as quasars.
The underlying energy source that drives the outflow is unknown. Promising
mechanisms invoke tapping the rotational energy stored in either the newly formed star
or the inner parts of its nebular disk. There exist theories suggesting that strong
magnetic fields coupled with rapid rotation act as whirling rotary blades to fling out the
nearby gas. Eventual collimation of the outflow toward the rotation axes appears to be a
generic feature of many proposed models.
Pre-main-sequence stars of low mass first appear as visible objects, T Tauri stars, with
sizes that are several times their ultimate main-sequence sizes. They subsequently
contract on a time scale of tens of millions of years, the main source of radiant energy in
this phase being the release of gravitational energy. As the internal temperature rises to
a few million kelvins, deuterium (heavy hydrogen) is first destroyed.
Then lithium, beryllium, and boron are broken down into helium as their nuclei are
bombarded by protons moving at increasingly high speeds. When their central
temperatures reach values comparable to 107 K, hydrogen fusion ignites in their cores,
and they settle down to long stable lives on the main sequence. The early evolution of
high-mass stars is similar; the only difference is that their faster overall evolution may
allow them to reach the main sequence while they are still enshrouded in the cocoon of
gas and dust from which they formed.
As the central temperature and density continue to rise, the proton-proton and carbon
cycles become active, and the development of the (now genuine) star is stabilized. The
star then reaches the main sequence, where it remains for most of its active life. The
time required for the contraction phase depends on the mass of the star. A star of
the Sun’s mass generally requires tens of millions of years to reach the main sequence,
whereas one of much greater mass might take a few hundred thousand years.
By the time the star reaches the main sequence, it is still chemically homogeneous.
With additional time, the hydrogen fuel in the core is converted to helium, and
the temperature slowly rises. If the star is sufficiently massive to have a convective core,
the matter in this region has a chance to be thoroughly mixed, but the outer region does
not mix with the core. The Sun, by contrast, has no convective core, and the helium-to-
hydrogen ratio is maximum at the center and decreases outward. Throughout the life of
the Sun, there has been a steady depletion of hydrogen, so that the concentration of
hydrogen at the center today is probably only about one-third of the original amount.
The rest has been transformed into helium. Like the rate of formation of a star, the
subsequent rate of evolution on the main sequence is proportional to the mass of the
star; the greater the mass, the more rapid the evolution. Whereas the Sun is destined to
endure for some 10 billion years, a star of twice the Sun’s mass burns its fuel at such a
rate that it lasts about 3 billion years, and a star of 10 times the Sun’s mass has a
lifetime measured in tens of millions of years. By contrast, stars having a fraction of the
mass of the Sun seem able to endure for trillions of years, which is much greater than
the current age of the universe.
Physical Science
The spread of luminosities and colors of stars within the main sequence can be
understood as a consequence of evolution. At the beginning of their lives as hydrogen-
burning objects, stars define a nearly unique line in the Hertzsprung-Russell
diagram called the zero-age main sequence. Without differences in initial
chemical composition or in rotational velocity, all the stars would start exactly from this
unique line. As the stars evolve, they adjust to the increase in the helium-to-hydrogen
ratio in their cores and gradually move away from the zero-age main sequence. When
the core fuel is exhausted, the internal structure of the star changes rapidly; it quickly
leaves the main sequence and moves toward the region of giants and supergiants.
As the composition of its interior changes, the star departs the main sequence slowly at
first and then more rapidly. When about 10 percent of the star’s mass has been
converted to helium, the structure of the star changes drastically. All of the hydrogen in
the core has been burned out, and this central region is composed almost entirely of
inert helium, with trace admixtures of heavier elements. The energy production now
occurs in a thin shell where hydrogen is consumed and more helium added to a growing
but inert core. The outer parts of the star expand outward because of the increased
burning there, and as the star swells up, its luminosity gradually increases. The details
of the evolutionary process depend on the metal-to-hydrogen ratio, and the course of
evolution differs for stars of different population types.
The great spread in luminosities and colors of giant, supergiant, and sub giant stars is
also understood to result from evolutionary events. When a star leaves the main
sequence, its future evolution is precisely determined by its mass, rate of rotation (or
angular momentum), and chemical composition and whether it is a member of a
close binary system. Giants and super giants of nearly the same radius and surface
temperature may have evolved from main-sequence stars of different ages and masses.
Theoretical calculations suggest that, as the star evolves from the main sequence,
the hydrogen-helium core gradually increases in mass but shrinks in size as more and
more helium ash is fed in through the outer hydrogen-burning shell. Energy is carried
outward from the shell by rapid convection currents. The temperature of the shell rises;
the star becomes more luminous; and it finally approaches the top of the giant domain
on the Hertzsprung-Russell diagram. By contrast, the core shrinks by gravitational
contraction, becoming hotter and denser until it reaches a central temperature of about
120 million K. At that temperature the previously inert helium is consumed in the
production of heavier elements.
When two helium nuclei each of mass 4 atomic units (4He) are jammed together, it
might be expected that they would form a nucleus of beryllium of mass 8 atomic units
(8Be). In symbols,4He + 4He → 8Be.Actually, however, 8Be is unstable and breaks down
into two helium nuclei. If the temperature and density are high enough, though, the
short-lived beryllium nucleus can (before it decays) capture another helium nucleus in
what is essentially a three-body collision to form a nucleus of carbon-12—namely,8Be
+ 4He → 12C.
This fusion of helium in the core, called the triple alpha process, can begin gradually in
some stars, but in stars with masses between about half of and three times the Sun’s
mass, it switches on with dramatic suddenness, a process known as the “helium flash.”
Outwardly the star shows no discernible effect, but the course of its evolution is
changed with this new source of energy. Having only recently become a red giant, it
now evolves somewhat down and then to the left in the Hertzsprung-Russell diagram,
becoming smaller and hotter. This stage of core helium burning, however, lasts only
about a hundredth of the time taken for core hydrogen burning. It continues until the
core helium supply is exhausted, after which helium fusion is limited to a shell around
the core, just as was the case for hydrogen in an earlier stage. This again sets the star
evolving toward the red giant stage along what is called the asymptotic giant branch,
located slightly above the main region of giants in the Hertzsprung-Russell diagram.
In more massive stars, this cycle of events can continue, with the stellar core reaching
ever-higher temperatures and fusing increasingly heavy nuclei, until the star eventually
experiences a supernova explosion (see below Evolution of high-mass stars). In lower-
mass stars like the Sun, however, there is insufficient mass to squeeze the core to the
temperatures needed for this chain of fusion processes to proceed, and eventually the
outermost layers extend so far from the source of nuclear burning that they cool to a few
thousand kelvins. The result is an object having two distinct parts: a well-defined core of
mostly carbon ash (a white dwarf star; see below End states of stars) and a swollen
spherical shell of cooler and thinner matter spread over a volume roughly the size of
the solar system. Such shells of matter, called planetary nebulas, are actually observed
in large numbers in the sky. Of the roughly 3,500 examples known in the Milky Way
Galaxy alone, NGC 7027 is the most intensively studied.
The relative abundances of the chemical elements provide significant clues regarding
their origin. Earth’s crust has been affected severely by erosion, fractionation, and
other geologic events, so that its present varied composition offers few clues as to its
early stages. The composition of the matter from which the solar system formed is
deduced from that of stony meteorites called chondrites and from the composition of
the Sun’s atmosphere, supplemented by data acquired from spectral observations of
hot stars and gaseous nebulas. The table lists the most abundant chemical elements; it
represents an average pertaining to all cosmic objects in general.
nitro 100,00
N Silicon Si 35,000 titanium Ti 91
gen 0
fluori manga
F 33 Sulfur S 17,400 Mn 288
ne nese
123,00
neon Ne chlorine Cl 250 iron Fe 33,000
0
sodiu
Na 2,100 Argon Ar 3,600 nickel Ni 1,800
m
The most obvious feature is that the light elements tend to be more abundant than the
heavier ones. That is to say, when abundance is plotted against atomic mass, the
resulting graph shows a decline with increasing atomic mass up to an atomic mass
value of about 100. Thereafter the abundance is more nearly constant. Furthermore, the
decline is not smooth. Among the lighter elements, those of even atomic number tend to
be more abundant, and those with an atomic number divisible by four are especially
favored. The abundances of lithium, beryllium, and boron are rare compared with those
of carbon, nitrogen, and oxygen. There is a pronounced abundance peak for iron and a
relatively high peak for lead, the most stable of the heavy elements.
The overwhelming preponderance of hydrogen suggests that all the nuclei were built
from this simplest element, a hypothesis first proposed many years ago and widely
accepted for a time. According to this now-defunct idea, all matter was initially
compressed into one huge ball of neutrons. As the universe began to expand,
its density decreased and the neutrons decayed into protons and electrons. The protons
then captured neutrons (see neutron capture), one after another, underwent beta
decay (ejection of electrons), and synthesized the heavy elements. A major difficulty
with this hypothesis, among various other problems, is that atomic masses 5 and 8 are
unstable, and there is no known way to build heavier nuclei by successive neutron
capture.
A large body of evidence now supports the idea that only the nuclei of hydrogen
and helium, with trace amounts of other light nuclei such as lithium, beryllium, and
boron, were produced in the aftermath of the big bang, the hot explosion from which the
universe is thought to have emerged, whereas the heavier nuclei were, and continue to
be, produced in stars. The majority of them, however, are fashioned only in the most
massive stars and some only for a short period of time after supernova explosions (see
below Evolution of high-mass stars).
Physical Science
The splitting in the spectral sequence among the cooler stars can be understood in
terms of composition differences. The M-type stars appear to have a normal (i.e., solar)
makeup, with oxygen more abundant than carbon and the zirconium group of elements
much less abundant than the titanium group. The R-type and N-type stars often contain
more carbon than oxygen, whereas the S-type stars appear to have
an enhanced content of zirconium as compared with titanium. Other
abundance anomalies are found in a peculiar class of higher temperature stars,
called Wolf-Rayet (or W) stars, in which objects containing predominantly helium,
carbon, and oxygen are distinguished from those containing helium and nitrogen, some
carbon, and little observed oxygen. Significantly, all these abundance anomalies are
found in stars thought to be well advanced in their evolutionary development. No main-
sequence dwarfs display such effects.
A most critical observation is the detection of the unstable element technetium in the S-
type stars. This element has been produced synthetically in nuclear laboratories on
Earth, and its longest-lived isotope, technetium-99, is known to have a half-life of
200,000 years. The implication is that this element must have been produced within the
past few hundred thousand years in the stars where it has been observed, suggesting
furthermore that this nucleosynthetic process is at work in at least some stars today.
This heavy element upwells from a star’s core (where it is produced) to the surface
(near where it is observed) in a phase called the third dredge-up, when material in deep
helium-burning layers is brought to the surface through convection.
Researchers have been able to demonstrate how elements might be created in stars by
nuclear processes occurring at very high temperatures and densities. No one
mechanism can account for all the elements; rather, several distinct processes
occurring at different epochs during the late evolution of a star have been proposed.
After hydrogen, helium is the most abundant element. Most of it was probably produced
in the initial big bang. Furthermore, as described earlier, helium is the normal ash of
hydrogen consumption, and in the dense cores of highly evolved stars, helium itself is
consumed to form, successively, carbon-12, oxygen-16, neon-20, and magnesium-24.
By this time in the core of a sufficiently massive star, the temperature has reached
some 700 million K. Under these conditions, particles such as protons, neutrons, and
helium-4 nuclei also can interact with the newly created nuclei to produce a variety of
other elements such as fluorine and sodium. Because these “uneven” elements are
produced in lesser quantities than those divisible by four, both the peaks and troughs in
the curve of cosmic abundances can be explained.
As the stellar core continues to shrink and the central temperature and density are
forced even higher, a fundamental difficulty is soon reached. A temperature of roughly
one billion K is sufficient to create silicon (silicon-28) by the usual method of helium
capture. This temperature, however, is also high enough to begin to break apart silicon
as well as some of the other newly synthesized nuclei. A “semi-equilibrium” is set up in
the star’s core—a balance of sorts between the production and destruction
(photodisintegration) of silicon. Ironically, though destructive, this situation is suitable for
the production of even heavier nuclei up to and including iron (iron-56), again through
the successive capture of helium nuclei.
If the temperature and the density of the core continue to rise, the iron-group nuclei tend
to break down into helium nuclei, but a large amount of energy is suddenly consumed in
the process. The star then suffers a violent implosion, or collapse, after which it soon
explodes as a supernova. In the catastrophic events leading to a supernova explosion
and for roughly 1,000 seconds thereafter, a great variety of nuclear reactions can take
place. These processes seem to be able to explain the trace abundances of all the
known elements heavier than iron.
Two situations have been envisioned, and both involve the capture of neutrons. When a
nucleus captures a neutron, its mass increases by one atomic unit and its charge
remains the same. Such a nucleus is often too heavy for its charge and might emit
an electron (beta particle) to attain a more stable state. It then becomes a nucleus of the
next higher element in the periodic table of the elements. In the first such process,
called the slow, or s-, process, the flux of neutrons is low. A nucleus captures a neutron
and leisurely emits a beta particle; its nuclear charge then increases by one.
Beta decay is often very slow, and, if the flux of neutrons is high, the nucleus might
capture another neutron before there is time for it to undergo decay. In this rapid, or r-,
process, the evolution of a nucleus can be very different from that in a slow process. In
supernova explosions, vast quantities of neutrons can be produced, and these could
result in the rapid buildup of massive elements. One interesting feature of the synthesis
of heavy elements by neutron capture at a high rate in a supernova explosion is that
nuclei much heavier than lead or even uranium can be fashioned. These in turn can
decay by fission, releasing additional amounts of energy.
The superabundant elements in the S-type stars come from the slow neutron process.
Moreover, the observation of technetium-99 is ample evidence that these processes are
at work in stars today. Even so, some low-abundance atomic nuclei are proton-rich (i.e.,
neutron-deficient) and cannot be produced by either the s- or the r-process.
Presumably, they have been created in relatively rare events—e.g., one in which
a quantum of hard radiation, a gamma-ray photon, causes a neutron to be ejected.
Finally, the peculiar A-type stars comprise a class of cosmic objects with strange
elemental abundance anomalies. These might arise from mechanical effects—for
example, selective radiation pressure or photospheric diffusion and element separation
—rather than from nuclear effects. Some stars show enhanced silicon, others
enhanced lanthanides. The so-called manganese stars show great overabundances of
manganese and gallium, usually accompanied by an excess of mercury. The latter stars
exhibit weak helium lines, low rotational velocities, and excess amounts of
gallium, strontium, yttrium, mercury, and platinum, as well as absences of such
elements as aluminum and nickel. When these types of stars are found in binaries, the
two members often display differing chemical compositions. It is most difficult
to envision plausible nuclear events that can account for the peculiarities of these
abundances, particularly the strange isotope ratios of mercury.
The final stages in the evolution of a star depend on its mass and angular
momentum and whether it is a member of a close binary.
White dwarfs
All stars seem to evolve through the red-giant phase to their ultimate state along a
straightforward path. In most instances, especially among low-mass stars, the distended
outer envelope of the star simply drifts off into space, while the core settles down as
a white dwarf. Here the star (really the core) evolves on the horizontal branch of
the Hertzsprung-Russell diagram to bluer colors and lower luminosities. In other cases,
in which the mass of the star is several solar masses or more, the star may explode as
a supernova. Even for these more massive stars, however, if the residual mass in the
core is less than 1.4 solar masses (the Chandrasekhar limit), the stellar remnant will
become a white dwarf. The matter in such a dwarf becomes a degenerate gas, wherein
the electrons are all stripped from their parent atoms. Gas in this peculiar state is an
almost perfect conductor of heat and does not obey the ordinary gas laws. It can be
compressed to very high densities, typical values being in the range of 10 million grams
per cubic cm (i.e., about 10 million times the density of water). Such a white dwarf no
longer has any source of energy and simply continues to cool down, eventually
becoming a black dwarf.
The energy output of a white dwarf is so small that the object can go on shining mainly
by radiating away its stored energy until virtually none is left to emit. How long this might
take is unknown, but it would seem to be on the order of trillions of years. The final
stage of this kind of low-mass star is typically a ball not much larger than Earth but with
a density perhaps 50,000 times that of water.
The Sun is destined to perish as a white dwarf. But, before that happens, it will evolve
into a red giant, engulfing Mercury and Venus in the process. At the same time, it will
blow away Earth’s atmosphere and boil its oceans, making the planet uninhabitable.
None of these events will come to pass for several billion years.
The first white dwarf to be recognized was the companion to Sirius. It was originally
detected by its gravitational attraction on the larger, brighter star and only later observed
visually as a faint object (now called Sirius B), about 10,000 times fainter than Sirius
(now called Sirius A) or 500 times fainter than the Sun. Its mass is slightly less than that
of the Sun, and its size a little less than that of Earth. Its color and spectrum correspond
roughly to spectral type A, with a surface temperature of about 25,000 K. Hence, the
energy emission per unit area from the surface must be much greater than that of the
Sun. Because Sirius B is so faint, its surface area and thus its volume must be very
small, and its average density is on the order of 100,000 times that of water.
Another well-known white dwarf, designated BD + 16°516, is paired with a much cooler
K0 V dwarf in an eclipsing system. The two stars, whose centers are separated by
Physical Science
2,092,000 km (about 1,300,000 miles), revolve around each other with a period of 12.5
hours. The white dwarf produces pronounced excitation and heating effects in the K-
type star’s atmosphere. The white dwarf’s mass is about 0.6 that of the Sun, but its
diameter is only 16,000 km (10,000 miles); hence, its density is about 650,000 times
that of water.
Neutron stars
When the mass of the remnant core lies between 1.4 and about 2 solar masses, it
apparently becomes a neutron star with a density more than a million times greater than
even that of a white dwarf. Having so much mass packed within a ball on the order of 20
km (12 miles) in diameter, a neutron star has a density that can reach that of nuclear
values, which is roughly 100 trillion (1014) times the average density of solar matter or of
water. Such a star is predicted to have a crystalline solid crust, wherein bare atomic
nuclei would be held in a lattice of rigidity and strength some 18 orders of magnitude
greater than that of steel. Below the crust, the density is similar to that of an atomic
nucleus, so the residual atomic cores lose their individuality as their nuclei are jammed
together to form a nuclear fluid.
Although neutron stars were predicted in the 1930s, it was not until the late 1960s that
observers accidentally discovered a radio source emitting weak pulses, each lasting
about 0.3 second with a remarkably constant period of approximately 1.337 seconds.
Other examples of such an object, dubbed a pulsar for “pulsating radio star,” were soon
found.
A large body of evidence now identifies pulsars as rotating magnetized neutron stars.
All the energy emitted in the pulses derives from a slowing of the star’s rotation, but only
a small fraction is released in the form of radio-frequency pulses. The rest goes into
pulses observed elsewhere in the electromagnetic spectrum and into cosmic rays, with
perhaps some going into the emission of gravitational energy, or gravity waves. For
example, the pulsar at the center of the Crab Nebula, the most well-known of
modern supernovas, has been observed not only at radio frequencies but also at optical
and X-ray frequencies, where it emits 100 and 10,000 times, respectively, as
much radiation as in the radio spectrum. The slowing of the pulsar’s spin also supplies
the energy needed to account for the non-thermal, or synchrotron, emission from the
Crab Nebula, which ranges from X-rays to gamma rays.
Crab pulsar
Crab pulsar (NP 0532), as observed by the Hubble Space Telescope. The pulsar is the
left member of a pair of stars near the center of the picture. Its energy fuels the glowing
center of the Crab Nebula.
Physical Science
Pulsar radiation is polarized, both linearly and circularly, and can be understood in
terms of a rotating star having a powerful magnetic field of a trillion gauss. (By
contrast, Earth’s magnetic field is about 0.5 gauss.) Various mechanisms have been
proposed whereby charged particles can be accelerated to velocities close to that
of light itself. Possibly most, if not all, galactic cosmic rays originate from supernovas
and remnant pulsars.
pulsar
A pulsar emits two beams of electromagnetic radiation along its magnetic axis. If the
magnetic axis is offset from the rotational axis, as the star rotates, the beams will sweep
out circular paths instead of remaining in one position. An observer in the path of such a
beam will thus detect a periodic pulse of radiation as the beam sweeps by.
Modern observations have recorded sudden changes in the rotation rates of pulsars.
The Vela pulsar, for instance, has abruptly increased its spin rate several times. Such a
period change or “glitch” can be explained if the pulsar altered its radius by about one
centimeter; this sudden shrinkage of the crust is sometimes called a “starquake.” Pulsar
phenomena apparently last much longer than the observable supernova remnants in
which they were born, since well more than 2,500 pulsars have been cataloged and
only a few are associated with well-known remnants. Even so, the statistics of pulsars
are likely to be observationally biased, since signals from pulsars at great distances in
the Galaxy become distorted by ionized regions of interstellar space.
Black holes
If the core remnant of a supernova exceeds about two solar masses, it continues to
contract. The gravitational field of the collapsing star is predicted to be so powerful that
neither matter nor light can escape it. The remnant then collapses to a black hole—
a singularity, or point of zero volume and infinite density hidden by an event horizon at a
distance called the Schwarzschild radius, or gravitational radius. Bodies crossing the
event horizon, or a beam of light directed at such an object, would seemingly just
disappear—pulled into a “bottomless pit.”
The existence of black holes is well established, both on a stellar scale, such as exists
in the binary system Cygnus X-1, and on a scale of millions or billions of solar masses
at the center of some galaxies, such as M87. (See quasar.)
Physical Science
Black hole at the center of the massive galaxy M87 about 55 million light-years from
Earth as imaged by the Event Horizon Telescope (EHT). The black hole is 6.5 billion
times more massive than the Sun. This image was the first direct visual evidence of a
supermassive black hole and its shadow. The ring is brighter on one side because the
black hole is rotating, and thus material on the side of the black hole turning toward
Earth has its emission boosted by the Doppler effects. The shadow of the black hole is
about five and a half times larger than the event horizon, the boundary marking the
black hole's limits, where the escape velocity is equal to the speed of light.
History
The periodic table and a natural number for each element
(from lanthanum to lutetium inclusive) must have 15 members—no fewer and no more
—which was far from obvious from known chemistry at that time.
Missing elements]
After Moseley's death in 1915, the atomic numbers of all known elements from
hydrogen to uranium (Z = 92) were examined by his method. There were seven
elements (with Z < 92) which were not found and therefore identified as still
undiscovered, corresponding to atomic numbers 43, 61, 72, 75, 85, 87 and 91. [5] From
1918 to 1947, all seven of these missing elements were discovered.[6] By this time, the
first four transuranium elements had also been discovered, so that the periodic table
was complete with no gaps as far as curium (Z = 96).
The proton and the idea of nuclear electrons
In 1915, the reason for nuclear charge being quantized in units of Z, which were now
recognized to be the same as the element number, was not understood. An old idea
called Prout's hypothesis had postulated that the elements were all made of residues (or
"protyles") of the lightest element hydrogen, which in the Bohr-Rutherford model had a
single electron and a nuclear charge of one. However, as early as 1907, Rutherford
and Thomas Royds had shown that alpha particles, which had a charge of +2, were the
nuclei of helium atoms, which had a mass four times that of hydrogen, not two times. If
Prout's hypothesis were true, something had to be neutralizing some of the charge of
the hydrogen nuclei present in the nuclei of heavier atoms.
In 1917, Rutherford succeeded in generating hydrogen nuclei from a nuclear
reaction between alpha particles and nitrogen gas,[7] and believed he had proven Prout's
law. He called the new heavy nuclear particles protons in 1920 (alternate names being
proutons and protyles). It had been immediately apparent from the work of Moseley that
the nuclei of heavy atoms have more than twice as much mass as would be expected
from their being made of hydrogen nuclei, and thus there was required a hypothesis for
the neutralization of the extra protons presumed present in all heavy nuclei. A helium
nucleus was presumed to be composed of four protons plus two "nuclear electrons"
(electrons bound inside the nucleus) to cancel two of the charges. At the other end of
the periodic table, a nucleus of gold with a mass 197 times that of hydrogen was
thought to contain 118 nuclear electrons in the nucleus to give it a residual charge of
+79, consistent with its atomic number.
The discovery of the neutron makes Z the proton number
All consideration of nuclear electrons ended with James Chadwick's discovery of the
neutron in 1932. An atom of gold now was seen as containing 118 neutrons rather than
118 nuclear electrons, and its positive charge now was realized to come entirely from a
content of 79 protons. After 1932, therefore, an element's atomic number Z was also
realized to be identical to the proton number of its nuclei.
The symbol of Z
The conventional symbol Z possibly comes from the German word Atomzahl (atomic
number).[8] However, prior to 1915, the word Zahl (simply number) was used for an
element's assigned number in the periodic table.
Chemical properties
Each element has a specific set of chemical properties as a consequence of the number
of electrons present in the neutral atom, which is Z (the atomic number).
The configuration of these electrons follows from the principles of quantum mechanics.
The number of electrons in each element's electron shells, particularly the
outermost valence shell, is the primary factor in determining its chemical
bonding behavior. Hence, it is the atomic number alone that determines the chemical
properties of an element; and it is for this reason that an element can be defined as
consisting of any mixture of atoms with a given atomic number.
New elements
The quest for new elements is usually described using atomic numbers. As of 2019, all
elements with atomic numbers 1 to 118 have been observed. Synthesis of new
Physical Science
Module 2
Week 3 and 4
MOST ESSENTIAL LEARNING COMPETENCIES
1. Determine if a molecules is polar or non-polar given its structure
2. Relate the polarity of a molecule to its properties
3. Describe the general types of inter-molecular forces
4. Explain the effect of intermolecular forces on the properties of substances
5. Explain how the structures of biological macromolecules such as carbohydrates,
lipids, nucleic acid, and proteins determine their properties and functions.
A molecule is a group of atoms. It is the smallest unit that can participate in a chemical
reaction.
There are numerous different types of molecules, and all of these can be categorized
into polar and non-polar groups. They are separated from each in the presence or
absence of electric poles. Let’s explore further:
Polar molecules have electric poles while non-polar molecules do not have electric
poles.
1. Polar molecules have positive and negative ends (charges) while non-polar
molecules don’t because their charges cancel out.
2. Distribution of electrons is equal in non-polar molecules. They are completely
symmetric. In polar molecules, we see that the charge is unevenly distributed.
3. When a highly electronegative atom bond with a relatively less Electronegativity
atom, a polar molecule is formed.
4. Most of the hydrocarbons liquids are non-polar.
5. Electrical dipole movement is seen in polar molecules and not in non-polar
molecules.
6. Polar molecules interact with other molecules of similar polarity to form solutions.
Non-polar molecules do not interact the same way.
7. Both types of molecules go by “like dissolves like” principle, which means that polar
molecules can dissolve into other polar molecules and non-polar into other non-polar
molecules. Polar cannot dissolve into non-polar molecules and vice versa.
Examples of polar and non-polar molecules
Polar molecules: Water, alcohol, Sulfur dioxide, ammonia, ethanol, hydrogen sulfide,
bent molecules (those with a significant bond angle) in general
Non-polar molecules: Hydrocarbons (gasoline, toluene), homo-nuclear diatomic
Physical Science
molecules (O2, N2, Cl2, H2, etc.), noble gases, benzene, methane, ethylene, carbon
tetrachloride
How to determine the polarity of molecules
Start by drawing its Lewis structure. This rule applies to all molecules except
hydrocarbons and molecules with two atoms of the same element.
The Lewis structure will help you analyze the shape of the molecule given to you.
Determine which of the five categories of shapes your molecule falls into linear,
tetrahedral, trigon planar, bent, trigon pyramid. The first three are symmetric
shapes, and the last two are asymmetric shapes.
As learned before, non-polar molecules are perfectly symmetrical while polar
molecules are not. This means that if the shape of the molecule given to you is a
bent or trigon pyramid, it is a polar molecule.
Remember that asymmetry applies even if the outer atoms are the same. The
arrangement of the atoms matters more.
Now, let’s dissect the symmetric molecules. All the atoms that are attached to the
central atom must be the same if it is a non-polar molecule. If different kinds of
atoms are attached to the central atom, the molecule is polar.
Are all bent molecules polar?
Mostly, yes. As aforesaid, bent molecules are asymmetrical just like trigon pyramids and
that means that they are polar molecules.
Examples of bent molecules are H2O, NO2, CH2, and SCl2.
The exceptions
There are a few exceptions to the rules of polar and non-polar molecules, and C-H bond
is a classic example. This molecule is non-polar even though the bonds are slightly
polar.
There are three main properties of chemical bonds that must be considered—namely,
their strength, length, and polarity. The polarity of a bond is the distribution of electrical
charge over the atoms joined by the bond. Specifically, it is found that, while bonds
between identical atoms (as in H2) are electrically uniform in the sense that
both hydrogen atoms are electrically neutral, bonds between atoms of different
elements are electrically in equivalent. In hydrogen chloride, for example, the
hydrogen atom is slightly positively charged whereas the chlorine atom is slightly
negatively charged. The slight electrical charges on dissimilar atoms are called partial
charges, and the presence of partial charges signifies the occurrence of a polar bond.
The existence of equal but opposite partial charges on the atoms at each end of a
hetero nuclear bond (i.e., a bond between atoms of different elements) gives rise to
an electric dipole. The magnitude of this dipole is expressed by the value of
its dipole moment, μ, which is defined as the product of the magnitude of the partial
charges times their separation (essentially, the length of the bond). The dipole moment
of a hetero nuclear bond can be estimated from the electronegativity of the atoms A and
B, χA and χB, respectively, by using the simple relation
Physical Science
where D denotes the unit debye, which is used for reporting molecular dipole moments
(1 D = 3.34 × 10−30 coulomb·meter). Moreover, the negative end of the dipole lies on the
more electronegative atom. If the two bonded atoms are identical, it follows that the
dipole moment is zero and the bond is nonpolar.
Even a homo nuclear bond, which is a bond between atoms of the same element (as
in Cl2), is not purely covalent, because a more accurate description would be in terms
of ionic-covalent resonance:
That the species is nonpolar despite the occurrence of ionic contributions stems from
the equal contributions of the ionic structures Cl−Cl+ and Cl+Cl− and their canceling
dipoles. That Cl2 is commonly regarded as a covalently bonded species stems from the
dominant contribution of the structure Cl―Cl to this resonance mixture. In contrast, the
valence bond theory wave function (see below Valence bond theory) of hydrogen
chloride would be expressed as the resonance hybrid
In this case, the two ionic structures would contribute different amounts (because the
elements have different electronegativity), and the larger contribution of H +Cl− is
responsible for the presence of partial charges on the atoms and the polarity of the
molecule.
A polyatomic molecule will have polar bonds if its atoms are not identical. However,
whether or not the molecule as a whole is polar (i.e., has a nonzero electric dipole
moment) depends on the shape of the molecule. For example, the carbon-oxygen
bonds in carbon dioxide are both polar, with the partial positive charge on the carbon
atom and the partial negative charge on the more electronegative oxygen atom. The
molecule as a whole is nonpolar, however, because the dipole moment of one carbon-
oxygen bond cancels the dipole moment of the other, for the two bond dipole moments
Physical Science
point in opposite directions in this linear molecule. In contrast, the water molecule is
polar. Each oxygen-hydrogen bond is polar, with the oxygen atom bearing the partial
negative charge and the hydrogen atom the partial positive charge. Because the
molecule is angular rather than linear, the bond dipole moments do not cancel, and the
molecule has a nonzero dipole moment.
The polarity of H2O is of profound importance for the properties of water. It is partly
responsible for the existence of water as a liquid at room temperature and for the ability
of water to act as a solvent for many ionic compounds. The latter ability stems from the
fact that the partial negative charge on the oxygen atom can emulate the negative
charge of anions that surround each cation in the solid and thus help minimize
the energy difference when the crystal dissolves. The partial positive charge on the
hydrogen atoms can likewise emulate that of the cation surrounding the anion in the
solid.
The preceding discussion has outlined the general approach to covalent bonding and
has shown how it is still widely employed for a qualitative understanding of molecules. It
is incomplete in many respects, however. First, the role of the electron pair remains
unexplained but appears to be the hinge of both Lewis’s theory and the VSEPR theory.
Second, there is evidence that suggests that Lewis’ theory overemphasizes the role of
electron pairs. More fundamentally, little has been said about the distribution of bonding
electrons in terms of orbitals, although it has been shown that in atoms the distributions
of electrons are described by wave functions. Finally, the models that have been
described have little quantitative content: they do not lead to bond lengths or precise
bond angles, nor do they give much information about the strengths of bonds.
A full theory of the chemical bond needs to return to the roots of the behaviour of
electrons in molecules. That is, the role of the electron pair and the quantitative
description of bonding must be based on the Schrödinger equation and the Pauli
exclusion principle. This section describes the general features of such an approach.
Once again, the discussion will be largely qualitative and conceptual rather than
mathematical and numerical. However, the character of the presentation here should
not be taken to imply that the current understanding of molecules is not rigorous,
quantitative, and precise.
Several difficulties are encountered at the outset of the application of the Schrödinger
equation to molecules. Even the simplest molecules consist of two nuclei and several
electrons, and interesting molecules may contain a thousand atoms and tens of
thousands of electrons. So that any progress of a generally applicable kind can be
made, approximations are necessary.
The depth of the minimum is (apart from a small correction for the vibrational properties
of the bond) equal to the bond dissociation energy and hence indicates the tightness
with which the two atoms are held together. The steepness of the walls of the curve,
which shows how rapidly the energy changes as the nuclear separation changes,
indicates the rigidity of the bond. Thus, quantitative information can be obtained from
such an approach.
A molecular potential energy curve. The strength of the bond is indicated by the
depth of the well below the energy of the separated atoms (to the right), and the bond
length is the corresponding inter nuclear separation.
The basis of VB theory is the Lewis concept of the electron-pair bond. Broadly
speaking, in VB theory a bond between atoms A and B is formed when two
atomic orbitals, one from each atom, merge with one another (the technical term
is overlap), and the electrons they contain pair up (so that their spins are ↓↑). The
merging of orbitals gives rise to constructive interference—i.e., an enhancement of
amplitude—between the wave functions in the areas where they overlap, and hence
an enhanced amplitude results in the inter nuclear region. As a consequence of the
formation of this region of heightened amplitude, there is an increased probability of
finding he electrons in the inter nuclear region (so echoing Lewis’s conception of the
bond) and, by implication, a lowering of the energy of the molecule.
The VB theory can be put in the broader context of quantum mechanics by drawing on
the superposition principle and the Pauli exclusion principle. The two principles
establish more precisely the type of orbital merging that is required and also show that,
Physical Science
to achieve that merging, the two electrons must pair their spins. The technical
justification will not be presented here.
As an illustration of the VB procedure, consider the structure of H2O. First, note that the
valence-shell electron configuration of an oxygen atom is 2s22px22py12pz1, with an
unpaired electron in each of two 2p orbitals, and
is the Lewis diagram for the atom. Each hydrogen atom has an unpaired 1s electron
(H·) that can pair with one of the unpaired oxygen 2p electrons. Hence, a bond can form
by the pairing of each hydrogen electron with an oxygen electron and the overlap of the
orbitals they occupy. The electron distribution arising from each overlap is cylindrically
symmetrical around the respective O―H axis and is called a σ bond. The VB
description of H2O is therefore that each hydrogen atom is linked to the oxygen atom by
a σ bond formed by pairing of a hydrogen 1s electron and an oxygen 2p electron.
Because a wave function can be written for this structure, an energy can be calculated
by solving the Schrödinger equation, and a bond length can be determined by varying
the nuclear separation and identifying the separation that results in the minimum
energy.
The term σ bond is widely used in chemistry to denote an electron distribution like that
in an oxygen-hydrogen bond, specifically one that has cylindrical symmetry about the
line between the two bonded atoms. It is not the only type of bond, however, as can be
appreciated by considering the structure of a nitrogen molecule, N2. Each nitrogen atom
has the valence-shell electron configuration 2s22px12py12pz1. If the z direction is taken to
lie along the inter nuclear axis of the molecule, then the electrons in the two
2pz orbitals can pair and overlap to form a σ bond. However, the 2px orbitals now lie in
the wrong orientation for head-to-head overlap, and they overlap side-to-side instead.
The resulting electron distribution is called a π bond. A π bond also helps to hold the
two atoms together, but, because the region of maximum electron density produced by
the overlap is off the line of the inter nuclear axis, it does not do so with the same
strength as a σ bond. The 2py electrons can pair and overlap in the same way and give
rise to a second π bond. Therefore, the structure of an N2 molecule consists of one σ
bond and two π bonds. Note how this corresponds to and refines the Lewis description
of the :N≡N: molecule.
Promotion of electrons
Valence bond theory runs into an apparent difficulty with CH4. The valence-shell
electron configuration of carbon is 2s22px12py1, which suggests that it can form only two
bonds to hydrogen atoms, in which case carbon would have a valence of 2. The normal
valence of carbon is 4, however. This difficulty is resolved by noting that only the overall
energy of a molecule is important, and, as long as a process leads to a lowering of
energy, it can contribute even if an initial investment of energy is required. In this case,
VB theory allows promotion to occur, in which an electron is elevated to a higher orbital.
Thus, a carbon atom is envisaged as undergoing promotion to the valence configuration
2s12px12py12pz1 as a CH4 molecule is formed. Although promotion requires energy, it
Physical Science
enables the formation of four bonds, and overall there is a lowering of energy. Carbon is
particularly suited to this promotion because the energy involved is not very great;
hence the formation of tetravalent carbon compounds is the rule rather than the
exception.
Hybridization
The discussion is not yet complete, however. If this description of carbon were taken at
face value, it would appear that, whereas three of the CH bonds in methane are formed
from carbon 2p orbitals, one is formed from a carbon 2s orbital. It is well established
experimentally, however, that all four bonds in methane are identical.
Ethylene
Figure 11: The shape of sp2 hybrid orbitals and the structure of an ethylene molecule.
The σ bond is formed by the overlap of hybrid atomic orbitals, and the π bond is formed
by the overlap of an hybridized p orbitals. The entire structure is resistant to twisting
around the carbon-carbon bond.
Physical Science
Resonant structures
The description of the planar hexagonal benzene molecule, C6H6, illustrates another
aspect of VB theory. Each of the six carbon atoms is taken to be sp2 hybridized. Two of
the hybrid orbitals are used to form σ bonds with the carbon atom neighbors, and one is
used to form a σ bond with a hydrogen atom. The unhybridized carbon 2p orbitals are in
a position to overlap and form π bonds with their neighbors. However, there are several
possibilities for pairing; two are as follows:
Benzene
Figure 12: The valence-bond description of a benzene molecule. The sp2 hybridized
carbon atoms form σ bonds with their neighbors, and the unhybridized p orbitals overlap
to form π bonds. This bonding pattern corresponds to one of the Kekulé structures (see
text).
There is a VB wave function for each of these so-called Kekulé structures. (They are so
called after Friedrich August Kekulé, who is commonly credited with having first
proposed the hexagonal structure for benzene in 1865; however, a cyclic structure had
already been proposed by Joseph Loschmidt four years earlier.) The actual structure is
a superposition (sum) of the two wave functions: in VB terms, the structure of benzene
is a resonance hybrid of the two canonical structures. In quantum mechanical terms, the
blending effect of resonance in the Lewis approach to bonding is the superposition of
wave functions for each contributing canonical structure. The effect of resonance is the
sharing of the double-bond character around the ring, so that each carbon-carbon bond
has a mixed single- and double-bond character. Resonance also
Physical Science
(for quantum mechanical reasons) lowers the energy of the molecule relative to either
contributing canonical structure. Indeed, benzene is a molecule that is surprisingly
resistant to chemical attack (double bonds, rather than being a source of molecular
strength and stability, are usually the targets of chemical attack) and is more stable than
its structure suggests.
One of the difficulties that has rendered VB computationally unattractive is the large
number of canonical structures, both covalent and ionic, that must be used in order to
achieve quantitatively reliable results; in some cases tens of thousands of structures
must be employed. Nevertheless, VB theory has influenced the language
of chemistry profoundly, and the concepts of σ and π bonds, hybridization, and
resonance are a part of the everyday vocabulary of the subject.
Just as an atomic orbital is a wave function that describes the distribution of an electron
around the nucleus of an atom, so a molecular orbital (an MO) is a wave function that
describes the distribution of an electron over all the nuclei of a molecule. If the
amplitude of the MO wave function is large in the vicinity of a particular atom, then the
electron has a high probability of being found there. If the MO wave function is zero in a
particular region, then the electron will not be found there.
The procedure can be introduced by considering the H2 molecule. Its molecular orbitals
are constructed from the valence-shell orbitals of each hydrogen atom, which are the
1s orbitals of the atoms. Two super positions of these two orbitals can be formed, one
by summing the orbitals and the other by taking their difference. In the former, the
amplitudes of the two atomic orbitals interfere constructively with one another, and there
is consequently an enhanced amplitude between the two nuclei. As a result, any
electron that occupies this molecular orbital has a high probability of being found
between the two nuclei, and its energy is lower than when it is confined to either atomic
orbital alone. This combination of atomic orbitals is therefore called a bonding orbital.
Moreover, because it has cylindrical symmetry about the inter nuclear axis, it is
designated a σ orbital and labeled 1σ.
The MO formed by taking the difference of the two 1s orbitals also has cylindrical
symmetry and hence is also a σ orbital. Taking the difference of the two atomic orbitals,
however, results in destructive interference in the inter nuclear region where the
amplitude of one orbital is subtracted from the other. This destructive interference is
Physical Science
complete on a plane midway between the nuclei, and hence there is a nodal plane—i.e.,
a plane of zero amplitude—between the nuclei. Any electron that occupies this orbital is
excluded from the inter nuclear region, and its energy is higher than it would be if it
occupied either atomic orbital. The orbital arising in this way is therefore called an anti-
bonding orbital; it is often denoted σ* (and referred to as “sigma star”) or, because it is
the second of the two σ orbitals, 2σ.
The molecular orbital energy-level diagram, which is a diagram that shows the relative
energies of molecular orbitals, for the H2 molecule is shown in Figure 13. On either side
of the central ladder are shown the energies of the 1s orbitals of atoms A and B, and the
central two-rung ladder shows the energies of the bonding and anti-bonding
combinations. Only at this stage, after setting up the energy-level diagram, are the
electrons introduced. In accord with the Pauli exclusion principle, at most two electrons
can occupy any one orbital. In H2 there are two electrons, and, following the building-up
principle, they enter and fill the lower-energy bonding combination. Hence the electron
configuration of the molecule is denoted 1σ2, and the stability of the molecule stems
from the occupation of the bonding combination. Its low energy results in turn (in the
conventional interpretation, at least) from the accumulation of electron density in the
inter nuclear region because of constructive interference between the contributing
atomic orbitals.
Figure 13: A molecular orbital energy-level diagram showing the relative energies of the
atomic orbitals of atoms A and B (1sA and 1sB) and the bonding (1σ) and antibonding
(2σ) molecular orbitals they form.
The central importance of the electron pair for bonding arises naturally in MO theory via
the Pauli exclusion principle. A single electron pair is the maximum number that can
occupy a bonding orbital and hence give the greatest lowering of energy. However, MO
theory goes beyond Lewis’s approach by not ascribing bonding to electron pairing;
some lowering of energy is also achieved if only one electron occupies a bonding
orbital, and so the fact that H2+ exists (with the electron configuration 1σ1) is no longer
puzzling.
The molecular orbital energy-level diagram shown in Figure 13 also applies (with
changes of detail in the energies of the molecular orbitals) to the hypothetical species
He2. However, this species has four valence electrons, and its configuration would be
1σ22σ2. Although there is a bonding influence from the two bonding electrons, there is
an anti-bonding influence from two anti-bonding electrons. As a result, the He 2 molecule
does not have a lower energy than two widely separated helium atoms and hence has
no tendency to form. (The overall effect is in fact slightly anti-bonding.) The role of
the noble gas configuration now can be seen from a different perspective: the electrons
that are provided by each closed-shell atom fill both the bonding and anti-bonding
orbitals, and they result in no net lowering of energy; in fact, they give rise to an
increase in energy relative to the separated atoms.
the electrons those orbitals contain). Then the sets of these orbitals that have the same
symmetry with respect to the inter nuclear axis are selected. (This point is illustrated
below.) Bonding and anti-bonding combinations of each set are then formed, and
from n atomic orbitals n such molecular orbitals are formed. The molecular orbital
energy- level diagram that results is constructed by putting the molecular orbitals in
order of increasing number of inter nuclear nodal planes, the orbital with no such nodal
plane lying at lowest energy and the orbital with nodal planes between all the atoms
lying at highest energy. At this stage, the valence electrons provided by the atoms are
allowed to occupy the available orbitals in accord with the general rules of the building-
up principle, with no more than two electrons in each orbital and in accord with Hund’s
rule if more than one orbital is available for occupation.
As a first illustration of this procedure, consider the structures of the diatomic molecules
formed by the period-2 elements (such as N2 and O2). Each valence shell has one
2s and three 2p orbitals, and so there are eight atomic orbitals in all and hence eight
molecular orbitals that can be formed. The energies of these atomic orbitals are shown
on either side of the molecular orbital energy-level diagram in Figure 14. (It may be
recalled from the discussion of atoms that the 2p orbitals have higher energy than the
2s orbitals.) If the z axis is identified with the inter nuclear axis, the 2s and 2pz orbitals
on each atom all have cylindrical symmetry around the axis and hence may be
combined to give σ orbitals. There are four such atomic orbitals, so four σ orbitals can
be formed. These four molecular orbitals lie typically at the energies shown in the
middle of Figure 14. The 2px orbitals on each atom do not have cylindrical symmetry
around the inter nuclear axis. They overlap to form bonding and anti-bonding π orbitals.
(The name and shape reflects the π bonds of VB theory.) The same is true of the
2py orbitals on each atom, which form a similar pair of bonding and anti-bonding π
orbitals whose energies are identical to those of bonding and anti-bonding π orbitals,
respectively, formed from the 2px orbitals. The precise locations of the π orbitals relative
to those of the σ orbitals depend on the species: for simplicity here they will be taken to
be as shown in Figure 14.
The molecular orbital energy-level diagram for diatomic molecules of period-2 elements.
The occupation of the orbitals is characteristic of N2.
the configuration 1σ22σ21π43σ2. Note that only the orbitals in the lower portion of the
diagram of Figure 14 are occupied. This configuration accounts for the considerable
strength of the bonding in N2 and consequently its ability to act as a diluent for
the oxygen in the atmosphere, because the O2 molecules are much more likely to react
than the N2 molecules upon collision with other molecules. An analysis of the identities
of the orbitals shows, after allowing for the cancellation of bonding effects by
antibonding effects, that the form of the electron configuration is (σ bonding orbital) 2(π
bonding orbitals)4. If each doubly occupied σ orbital is identified with a σ bond and each
doubly occupied π orbital with a π bond, then the structure obtained by this MO
procedure matches both the VB description of the molecule and the :N≡N: Lewis
description.
To see how the MO approach transcends the Lewis approach (and, in this instance, the
VB approach as well), consider the electronic configuration of O2. The same MO
energy-level diagram (with changes of detail) can be used because the oxygen atoms
provide the same set of atomic orbitals. Now, however, there are 2 × 6 = 12 valence
orbitals to accommodate. The first 10 electrons reproduce the configuration of N 2. The
last two enter the 2π* anti-bonding orbital, thereby reducing the net configuration to one
σ bond and one π bond. That is, O2 is a doubly-bonded species, in accord with the
Lewis structure O=O. However, because there are two 2π orbitals and only two
electrons to occupy them, the two electrons occupy different orbitals with parallel spins
(recall Hund’s rule). Therefore, the magnetic fields produced by the two electrons do not
cancel, and O2 is predicted to be a paramagnetic species. That is in fact the case. Such
a property was completely outside the competence of Lewis’s theory to predict and
must be contrived in VB theory. It was an early major triumph of MO theory.
The principal qualitative difference between MO theory and VB theory becomes obvious
when the objects of study are polyatomic, rather than diatomic, species.
The benzene molecule is considered again but in this case from the viewpoint of its
molecular orbitals. The atomic orbitals that provide the so-called basis set for the
molecular orbitals (i.e., those from which the MOs are constructed) are
the carbon 2s and 2p orbitals and the hydrogen 1s orbitals. All these orbitals except one
2p orbital on each carbon atom lie in the plane of the molecule, so they naturally form
two sets that are distinguished by their symmetries. This discussion concentrates on the
molecular orbitals that are constructed from the six perpendicular 2p orbitals, which
form the π orbitals of the molecule; the remaining orbitals form a framework of σ
orbitals.
Six molecular orbitals, which are labeled 1a, 1e, 2e, and 2a, as shown in Figure 15, can
be built from these six 2p orbitals. The two 1e orbitals and the two 2e orbitals each have
the same energy. The six molecular orbitals are various sums and differences of the six
2p orbitals, and they differ in the number and position of their internuclear nodal
planes (i.e., areas of low electron density). As before, the greater the number of these
nodal planes, the more the electrons that occupy the orbitals are excluded from the
region between the nuclei, and hence the higher the energy. The resulting molecular
orbital energy-level diagram is shown alongside the orbitals in the illustration. The
lowest-energy 1a orbital has no nodal plane, so there is maximum positive overlap. The
two degenerate 1e orbitals each have one nodal plane, the degenerate 2e orbitals have
two nodal planes each, and the high-energy 2a orbital has three nodal planes. The
crucial difference from the cases considered earlier is that the molecular orbitals spread
over more than two atoms. That is, they are delocalized orbitals, and electrons that
occupy them are delocalized over several atoms (here, as many as six atoms, as in the
1a orbital).
Physical Science
Benzene
Figure 15: The six π molecular orbitals of a benzene molecule and their relative
energies. Only the three lowest-energy orbitals are occupied in benzene. The bonding
and anti-bonding character of these orbitals is distributed around the ring of carbon
atoms. The dashed lines represent nodal planes, and the shading reflects the two
possible phases of the orbitals. Constructive interference, resulting in an area of high
electron density, occurs between like phases; destructive interference, resulting in a
nodal plane, occurs between unlike phases.
Each carbon atom supplies one electron to the π system (the other 24 valence
electrons have occupied the 12 low-energy σ orbitals that are not directly of interest
here). These six electrons occupy the three lowest-energy molecular orbitals. Notice
that none of the net anti-bonding orbitals is occupied. This is a part of the explanation of
the considerable stability of the benzene molecule.
In the VB description of the benzene molecule, each double bond is localized between a
particular pair of atoms, but resonance spreads that character around the ring. In MO
theory, there are three occupied π orbitals, and hence three contributions to double-
bond character, but each electron pair is spread around the ring and helps to draw
either all the atoms together (the 1a orbital) or several of the atoms together (the two
1e orbitals). Thus, delocalization distributes the bonding effect of an electron pair over
the atoms of the molecule, and hence one electron pair can contribute to the bonding of
more than two atoms.
Several problems that remained unsolved in the earlier discussion of Lewis structures
can be unraveled. It has already been shown that one electron can contribute to
bonding if it occupies a bonding orbital; therefore the problem of the existence of one-
electron species is resolved.
Hypervalence is taken care of, without having to invoke octet expansion, by the
distributed bonding effect of delocalized electrons. Consider SF6, which according
to Lewis’s theory needs to use two of its 3d orbitals in addition to its four 3s and
3p orbitals to accommodate six pairs of bonding electrons. In MO theory, the four
3s and 3p orbitals of sulfur and one 2p orbital of each fluorine atom are used to build 1
+ 3 + 6 = 10 molecular orbitals. These 10 MOs are delocalized to varying degrees over
the seven atoms of the molecule. Half of them have a net bonding character and half of
Physical Science
them a net anti-bonding character between the sulfur and fluorine atoms. There are 6
sulfur valence electrons to accommodate and 6 × 1 = 6 fluorine electrons for a total of
12. The first 10 of these electrons occupy the net bonding orbitals; the remaining two
occupy the lowest-energy anti-bonding orbital. In fact, this orbital is so weakly anti-
bonding that it is best to regard it as nonbonding and as having little effect on the
stability of the molecule. In any event, its weakly anti-bonding character is distributed
over all six fluorine atoms, just as the other five pairs of electrons help to bind all six
fluorine atoms to the central sulfur atom. The net effect of the 12 electrons is therefore
bonding, and delocalization eliminates the need to invoke any role for d orbitals. The
quantitative description of the forms and energies of the molecular orbitals is improved
by the inclusion of 3d orbitals in the basis set, but only a small admixture is needed.
There is certainly no need to invoke 3d orbitals as a necessary component of the
description of bonding and no need to regard this hypervalent molecule as an example
of a species with an expanded octet. Octet expansion is a rule of thumb, a correlation of
an observation with the presence of available d orbitals, and not a valid explanation.
The language that molecular orbital theory brings to chemistry is that of bonding and
anti bonding orbitals and delocalization of electrons. The theory is presented here as
an alternative to valence bond theory, and the formulation of the theory is quite different.
However, both theories involve approximations to the actual electronic structures of
molecules, and both can be improved. Valence bond theory is improved by
incorporating extensive ionic-covalent resonance; molecular orbital theory
is enhanced by allowing for a variety of occupation schemes for molecular orbitals (the
procedure of configuration interaction). As these two improvement schemes are
pursued, the wave functions generated by the two approaches converge on one another
and the electron distributions they predict become identical.
Valence bond theory is widely used when the molecular property of interest is
identifiable with the properties of individual bonds. It is therefore commonly employed
in organic chemistry, where the reactions of molecules are often discussed in terms of
the properties of their functional groups. The latter are small localized regions of a
molecule (such as a double bond) or particular clusters of atoms (such as an OH
group). Molecular orbital theory is widely used to describe properties that are most
naturally discussed in terms of delocalization. Such properties include the spectroscopic
properties of molecules, in which electromagnetic radiation is used to excite an electron
from one molecular orbital to another and all the atoms contribute to the shift in electron
density that accompanies the excitation.
Intermolecular Forces
Physical Science
Molecules cohere even though their ability to form chemical bonds has been satisfied.
The evidence for the existence of these weak intermolecular forces is the fact
that gases can be liquefied, that ordinary liquids exist and need a considerable input of
energy for vaporization to a gas of independent molecules, and that many
molecular compounds occur as solids. The role of weak intermolecular forces in the
properties of gases was first examined theoretically by the Dutch scientist Johannes van
der Waals, and the term van der Waals forces is used synonymously with
intermolecular forces. Under certain conditions, weakly bonded clusters of molecules
(such as an argon atom in association with a hydrogen chloride molecule) can exist;
such delicately bonded species are called van der Waals molecules.
There are many types of intermolecular forces; the repulsive force and four varieties of
attractive force are discussed here. In general, the energy of interaction varies with
distance, as shown by the graph in Figure 16. Attractive forces dominate to the distance
at which the two molecules come into contact, then strong repulsive forces come into
play and the potential energy of two molecules rises abruptly. The shape of the
intermolecular potential energy curve shown in the illustration resembles that of the
molecular potential energy curve in Figure 10. The minimum of the former is much
shallower, however, showing that forces between molecules are typically much weaker
than the forces responsible for chemical bonds within molecules.
An intermolecular potential energy curve. The graph shows how the potential energy of
two molecules varies with their separation. The energy minimum is shallower than for
the formation of a chemical bond between two atoms.
Repulsive force
Dipole–dipole interaction
The first of the four bonding interactions discussed here is the dipole–dipole interaction
between polar molecules. It will be recalled that a polar molecule has an electric
dipole moment by virtue of the existence of partial charges on its atoms. Opposite
partial charges attract one another, and, if two polar molecules are orientated so that
the opposite partial charges on the molecules are closer together than their like
charges, then there will be a net attraction between the two molecules. This type of
intermolecular force contributes to the condensation of hydrogen chloride to a liquid at
low temperatures. The dipole–dipole interaction also contributes to the weak
interaction between molecules in gases, because, although molecules rotate, they tend
to linger in relative orientations in which they have low energy—namely, the mutual
orientation with opposite partial charges close to one another.
Dipole–induced-dipole interaction
Dispersion interaction
The third type of interaction acts between all types of molecule, polar or not. It is also
somewhat stronger than the two attractive interactions discussed thus far and is the
principal force responsible for the existence of the condensed phases of certain
molecular substances, such as benzene, other hydrocarbons, bromine, and
the solid elements phosphorus (which consists of tetrahedral P4 molecules)
and sulfur (which consists of crown-shaped S8 molecules). The interaction is called the
dispersion interaction or, less commonly but more revealingly, the induced-dipole–
induced-dipole interaction. Consider two nonpolar molecules near each other. Although
there are no permanent partial charges on either molecule, the electron density can be
thought of as ceaselessly fluctuating. As a result of these fluctuations, regions of equal
and opposite partial charge arise in one of the molecules and give rise to a transient
dipole. This transient dipole can induce a dipole in the neighboring molecule, which then
interacts with the original transient dipole. Although the latter continuously flickers from
one direction to another (with an average of zero dipole overall), the induced dipole
follows it, and the two correlated dipoles interact favorably with one another and cohere.
bond is about 10 times as strong as the other interactions described above, and when
present it dominates all other types of intermolecular interaction. It is responsible, for
example, for the existence of water as a liquid at normal temperatures; because of its
low molar mass, water would be expected to be a gas. The hydrogen bond is also
responsible for the existence as solids of many organic molecules containing hydroxyl
groups (―OH); the sugars glucose and sucrose are examples.
Many interpretations of the hydrogen bond have been proposed. One that fits into the
general scheme of this article is to think of the A―H unit as being composed of an
A atomic orbital and a hydrogen 1s orbital and to consider a lone pair of electrons on B
as occupying a B orbital. When the three atoms are aligned, these three orbitals can
form three molecular orbitals: one bonding, one largely nonbonding, and one anti
bonding. There are four electrons to accommodate (two from the original A―H bond
and two from the lone pair). They occupy the bonding and nonbonding orbitals, leaving
the ant bonding orbital vacant. Hence, the net effect is to lower the energy of the AHB
grouping and thus to constitute an intermolecular bond. Once again, on encountering
the hydrogen bond, one encounters a twist in the conventional attitude; the question
raised by this interpretation is not why such a bond occurs but why it does not occur
more generally. The explanation lies in the small size of the hydrogen atom, which
enables the balance of energies in the molecular orbital scheme to be favorable to
bonding.
Hydrogen bonding occurs to atoms other than nitrogen, oxygen, and fluorine if they
carry a negative charge and hence are rich in readily available electrons.
Thus, hydrogen bonding is one of the principal mechanisms of hydration of anions in
aqueous solution (the bonding of H2O molecules to the solute species) and hence
contributes to the ability of water to act as a good solvent for ionic compounds. It also
contributes to the hydration of organic compounds containing oxygen or nitrogen atoms
and thus accounts for the much greater aqueous solubility of alcohols than
hydrocarbons.
Hydrogen bond
Figure 17: The linking of atoms in two peptide links by the hydrogen bonds they can
form. The links may be part of the same polypeptide chain that has doubled back on
itself, or they may belong to different chains.
Hydrogen bonds are also responsible for the transmission of genetic information from
one generation to another, for they are responsible for the specific keying together
of cytosine with guanine and thymine with adenine moieties that characterizes the
structure of the DNA double helix.
Varieties of Solids
Chemical bonds and intermolecular forces are jointly responsible for the existence of
the solid phases of matter. This section reviews some of the types of solid that are
encountered and relates them to the topics discussed earlier.
Ionic solids
The structures of ionic solids have already been described in some detail. They consist
of individual ions that are stacked together in such a way that the assembly has the
lowest possible energy. These ions may be monatomic (as in sodium chloride, which
consists of Na+ and Cl− ions) or the ions may themselves be covalently bonded
polyatomic species. An example of the latter is ammonium nitrate, in which the cation is
NH4+ and the anion is NO3−; the N―H and N―O bonds within the ions are covalent.
Ionic compounds are generally hard and brittle and have high melting points.
Molecular solids
The structures of molecular solids, which are solids composed of individual molecules,
have also been touched on in the section on intermolecular forces. These molecules are
held to one another by hydrogen bonds (if they can form them), dispersion forces, and
other dipolar forces—in that order of decreasing importance—and the molecules stack
together in a pattern that minimizes their total energy. Examples of such solids
include ice, in which hydrogen bonding is of paramount importance, and polyethylene,
Physical Science
in which dispersion forces are dominant. Unless hydrogen bonds are present (in which
case molecular solids resemble ionic solids in brittleness), molecular solids are
generally soft and have low melting points because the bonds between the molecules
are easily overcome.
Network solids
There exists a class of solids called network solids in which the bonding is essentially
due to a network of covalent bonds that extends throughout the solid. Such solids are
hard and rigid and have high melting points because the crystal is like one
enormous molecule. The most well-known example of a network solid is diamond,
which consists of tetrahedral bonded carbon atoms (see Figure 7). By virtue of the
rigidity of its bonding structure, diamond is the hardest substance known and also the
best conductor of heat.
Some solids have a network character in certain directions and a more molecular
character in other directions. Once again, carbon provides the paradigm example, for
the form of carbon known as graphite consists of a stack of sheets of hexagonal rings of
carbon atoms. In the plane of the sheets, the bonding is covalent (and resembles an
extended version of the bonding in benzene). The sheets themselves are held together
by binding that is so weak that it is sometimes referred to as a van der Waals
interaction. The anisotropy of the structure of graphite accounts for the anisotropy of its
electrical conductivity (which is higher in the plane of the sheets than perpendicular to
them). The ability of graphite to shed sheets of carbon (a feature utilized in the
manufacture of pencils) and to act as a high-temperature lubricant (because the sheets
can slide over one another) appears to be consistent with this structure but in fact
seems to depend on the presence of impurities between the sheets.
Metals
The remaining major type of solid is a metal. A metal is characterized by its lustre, the
ease with which it may be deformed (rather than shattered) by hammering, and its high
electrical and thermal conductivities. Metals also tend to have higher densities than
other types of solid. The starting point for theories of the structures of metals is to
regard them as consisting of cations of the metal atoms embedded in a sea formed by
the discarded valence electrons. The mobility of these electrons accounts for the
mechanical, optical, and electrical properties of metals. The spherical cations can pack
closely together yet still give rise to locally neutral electrical assemblies. This is because
of the ability of the electrons to spread between the cations and neutralize their charges
regardless of how closely they are packed. The closeness of the packing of the atoms
accounts for the high densities of metals.
In the context of theories of the chemical bond, a metal is one extremely large homo
nuclear molecule. (For an alternative point of view, see crystal.) If a sample
of sodium metal is thought of as consisting of n sodium atoms where each atom has a
3s orbital for use in the construction of molecular orbitals and each atom supplies
one electron to a common pool, then from these n atomic orbitals n molecular orbitals
can be constructed. Each orbital has a characteristic energy, and the range of energies
spanned by the n orbitals is finite, however great the value of n. If n is very large, it
follows that the energy separation between neighboring molecular orbitals is very small
and approaches zero as n approaches infinity. The molecular orbitals then form
a band of energies. Another similar band can be formed by the overlap of the 3p orbitals
of the atoms, but there is a substantial band gap—i.e., a region of energy in which there
are no molecular orbitals—between the two bands.
the uppermost filled orbitals, and it is easy for a perturbation, such as an applied
potential difference or an oscillating electromagnetic field of incident light, to move the
electrons into these unoccupied levels. Hence, the electrons are very mobile and can
conduct an electric current, reflect light, transmit energy, and rapidly migrate to new
locations when the cations are moved by hammering.
The full theory of the structure of metals is a highly technical subject (as are the full
theories of the other topics discussed here). This brief introduction has been intended
only to show that the ideas of molecular orbital theory can be naturally extended to
account for the general features of the structures and properties of solids.
This section treats several aspects of molecular structure that are of more specialized
interest and shows how particular classes of compounds are described. Molecular
orbital theory will be used as a framework for the discussion, but aspects
of valence bond theory will be incorporated when it is natural (in the sense of being
commonplace in chemistry) to use them.
A particular class of compounds that once gave rise to some difficulty in the explanation
of the origin of their bonding are the complexes of transition metal ions. There are
numerous examples of such species; they have in common a structure in which a
central metal ion is surrounded by a number of ions or molecules, called ligands, that
can also exist separately. The most common complexes have six ligands arranged in
an octahedron around the central ion. An example is [Fe(H2O)6]2+, where Fe
denotes iron. This species can essentially be regarded as an Fe2+ ion, with an electron
configuration [Ar]3d6, surrounded by six H2O molecules linked to the metal ion through
their oxygen atoms.
five d orbitals have the same energy, in an octahedral crystal field they split into two
groups (Figure 18), with three orbitals (labeled t2g; the labeling is based on details of
their symmetry) lower in energy than the remaining two (labeled eg). The difference in
energy between the two sets is denoted Δ and is called the crystal field splitting
energy (CFSE). This energy is the parameter that is used to correlate a variety
of spectroscopic, thermodynamic, and magnetic properties of complexes.
Figure 18: Crystal field splitting. In an octahedral complex, the d orbitals of the central
metal ion divide into two sets of different energies. The separation in energy is the
crystal field splitting energy, Δ. (A) When Δ is large, it is energetically more favourable
for electrons to occupy the lower set of orbitals. (B) When Δ is small, it is energetically
more favorable for the electrons to occupy both sets with as many parallel electron
spins as possible.
The essential feature of crystal field theory is that there is a competition between the
magnitude of the CFSE and the pairing energy, which is the energy required to
accommodate two electrons in one orbital. When the pairing energy is high compared
with the CFSE, the lowest-energy electron configuration is achieved with as many
electrons as possible in different orbitals. The arrangement of a d5 ion, for instance,
is t2g3eg2, with all spins parallel (as in Figure 18B). However, if the ligands give rise to a
very strong crystal field, so that the CFSE is large compared with the pairing energy,
then the lowest-energy electron configuration is that with as many electrons as possible
in the lower (t2g) set of orbitals. In such a case, a 3d5 ion would adopt the
configuration t2g5, with only one unpaired spin as in Figure 18A. Thus,
because magnetism arises from the presence of electron spins, it can be seen that the
magnetic properties of the complex can be correlated with the size of the CFSE. The
same is true of spectroscopic and thermodynamic properties. In particular it is found
that ligands can be arranged in order of the strength of the crystal field that they
generate, and this so-called spectrochemical series can be used to rationalize and
predict the properties of complexes.
cation and one orbital from each of the six ligand atoms that are directly attached to the
metal cation. It follows that in such an octahedral complex there are 5 + 6 = 11
molecular orbitals to accommodate the 3d electrons of an [Ar]3dn species and 12
electrons from the six ligand atoms, giving 12 + n electrons in all. The 11 MOs span a
range of energies. Twelve of the electrons occupy the six lowest-energy MOs, which are
largely ligand-atom in character. The remaining n electrons are to be accommodated in
the eg and t2g sets of orbitals. The energy separation between these two sets of orbitals,
the ligand-field splitting energy (LFSE) is the ligand field version of the CFSE in crystal
field theory, and from this point on the construction of the lowest-energy electron
configuration is much the same as in crystal field theory. However, ligand field theory is
less artificial, allows for electron delocalization, and is more readily extended to more
complex patterns of bonding between the central metal ion and the ligands (such as the
incorporation of bonds with π symmetry).
Organometallic compounds
The stabilities of organometallic compounds follow certain empirical rules, among which
the 18-electron rule is the analogue of the octet rule of main-group compounds.
According to this rule, the most stable organometallic compounds are those having 18
electrons in the valence shell, a term in this context extended to include the
outermost d orbitals. Nickel tetracarbonyl, Ni(CO)4, a poisonous gas used in the refining
of nickel, has 10 electrons provided by the neutral nickel atom and two from each of the
four CO ligands, giving 18 electrons in all.
Boranes
Diborane
Figure 19: The structure of the three-center, two-electron bond in a B―H―B fragment
of a diborane molecule. A pair of electrons in the bonding combination pulls all three
atoms together.
The usefulness of the concept of a 3c,2e bond stems from two observations. The first is
that diborane is in fact only one of a large class of compounds of boron and hydrogen,
the boranes and the borohydride anions, in which the same feature is found. The
second observation is that a 3c,2e bond can be formed by three boron atoms. Intricate
networks of atoms can be formed in this way—for example, some having the form of
closed frameworks (the closo-boranes), some looking like untidy birds’ nests (the nido-
boranes), and some resembling spiderwebs (the arachno-boranes). Which type of
structure is obtained correlates with the number of valence electrons in the molecule,
and the correlation is expressed by Wade’s rules. These rules are empirical, but they
can be justified by a consideration of the numbers of 3c,2e and ordinary 2c,2e bonds
that are needed in each type of structure. They constitute an excellent example of how
chemists utilize the concept of bond formation and deploy a mixture of valence bond
and molecular orbital concepts to establish or rationalize helpful correlations between
the number of electrons present and the structure of the species.
A metal cluster compound is one in which metal atoms are linked directly to one another
(Figure 20). A simple example is the ion Hg22+, in which two mercury (Hg) ions are
linked together. A slightly more elaborate version is the ion [Re2Cl8]2−, in which there is a
direct link between two rhenium (Re) atoms. Some metal cluster compounds have more
than two metal atoms; an example is [Re3Cl12]3−, in which there are three rhenium atoms
bonded together. It is sometimes difficult to determine whether the metal atoms are
indeed directly linked or merely held quite close together by a framework of
bridging ligands.
Physical Science
The [Re2Cl8]2− ion, with a metal-metal link that has quadruple-bond character.
Metal cluster compounds warrant a special mention here because they provide the only
examples of quadruple bonds in chemistry. Apart from that, their bonding can be treated
as a straightforward exercise in MO or VB theory. Indeed, a metal cluster can be
regarded as an exceedingly tiny sample of metal, with insufficient atoms present for the
molecular orbitals to form a continuous band. The structure of [Re2Cl8]2− is shown
in Figure 20. The clue to the existence of unusual bonding is the arrangement of the two
sets of chloride ligands: to minimize repulsions between the atoms, each ReCl4 group
might be expected to be twisted 45° relative to the next rather than being in the
orientation shown. There appears to be a bonding feature between the two rhenium
atoms that holds the groups as illustrated. This feature is taken to be a quadruple bond
arising from the overlap of d orbitals on the two rhenium atoms.
There are two strands of approach to the computation of molecular structure. In the
semi-empirical approach, the calculation draws on a number of experimentally
determined characteristics to help in the overall calculation. In the ab initio approach,
the calculation proceeds from first principles (the Schrödinger equation) and makes no
use of imported information. The former approach was dominant in the 1970s, but
increases in computing power have led to an ascendancy of ab initio techniques since
then. The latter are intrinsically more reliable because there can be no certainty that a
quantity determined in one context is appropriate to a particular molecule.
Thus, a particular electronic distribution is proposed, and the distribution of the electrons
is recalculated on the basis of this first approximation. The distribution is then calculated
again on the basis of that improved description, and the process is continued until there
is negligible change—i.e., until the electron distribution has achieved self-consistency.
The implementation of this basic strategy can take a number of forms, and rival
techniques have given rise to a large number of acronyms, such as AM1 (Austin
Method 1) and MINDO (Modified Intermediate Neglect of Differential Overlap), which
are two popular semi-empirical procedures.
With self-consistency established, the wave functions are available for detailed scrutiny.
One illustration must suffice. There is certain evidence
that carcinogenic or pharmacological activity correlates with certain aspects of
the charge distribution in molecules. Instead of dealing with the primitive concept of
partial charges, numerical wave functions can be used to map the details of the charge
distribution and hence to screen molecules for possible activity. This approach is
potentially of considerable utility for pharmaceutical products as it can help to reduce
the amount of in vivo screening of novel products.
The body cells then use dehydration reaction to assemble the monomers into new
polymers that carry out functions specific to the particular cell type.
Humans and other vertebrates store a day’s supply of glycogen in the liver and
muscles.
Cellulose is a major component of the tough wall of plant cells.
Plants produce almost one hundred billion tons of cellulose per year. It is the most
abundant organic compound on Earth.
Like starch, cellulose is a polymer of glucose. However, the glycosidic linkages in these
two polymers differ.
The difference is based on the fact that there are actually two slightly different ring
structures for glucose.
These two ring forms differ in whether the hydroxyl group attached to the number 1
carbon is fixed above (beta glucose) or below (alpha glucose) the plane of the ring.
Starch is a polysaccharide of alpha glucose monomers.
Cellulose is a polysaccharide of beta glucose monomers, making every other glucose
monomer upside down with respect to its neighbors.
The differing glycosidic links in starch and cellulose give the two molecules distinct
three-dimensional shapes.
While polymers built with alpha glucose form helical structures, polymers built with beta
glucose form straight structures.
The straight structures built with beta glucose allow H atoms on one strand to form
hydrogen bonds with OH groups on other strands.
In plant cell walls, parallel cellulose molecules held together in this way are grouped into
units called micro fibrils, which form strong building materials for plants (and for
humans, as lumber).
The enzymes that digest starch by hydrolyzing its alpha linkages cannot hydrolyze the
beta linkages in cellulose.
Cellulose in human food passes through the digestive tract and is eliminated in feces as
“insoluble fiber.”
As it travels through the digestive tract, cellulose abrades the intestinal walls and
stimulates the secretion of mucus, aiding in the passage of food.
Some microbes can digest cellulose to its glucose monomers through the use of
cellulase enzymes.
Many eukaryotic herbivores, from cows to termites, have symbiotic relationships with
cellulolytic microbes, providing the microbe and the host animal access to a rich source
of energy.
Some fungi can also digest cellulose.
Another important structural polysaccharide is chitin, used in the exoskeletons of
arthropods (including insects, spiders, and crustaceans).
Chitin is similar to cellulose, except that it contains a nitrogen-containing appendage on
each glucose monomer.
Pure chitin is leathery but can be hardened by the addition of calcium carbonate.
Chitin also provides structural support for the cell walls of many fungi.
Cholesterol is also the precursor from which all other steroids are synthesized.
Many of these other steroids are hormones, including the vertebrate sex hormones.
While cholesterol is an essential molecule in animals, high levels of cholesterol in the
blood may contribute to cardiovascular disease.
Both saturated fats and trans fats exert their negative impact on health by affecting
cholesterol levels.
Amino acids are the monomers from which proteins are constructed.
Amino acids are organic molecules with both carboxyl and amino groups.
At the center of an amino acid is an asymmetric carbon atom called the alpha carbon.
Four components are attached to the alpha carbon: a hydrogen atom, a carboxyl group,
an amino group, and a variable R group (or side chain).
Different R groups characterize the 20 different amino acids.
R groups may be as simple as a hydrogen atom (as in the amino acid glycine), or it may
be a carbon skeleton with various functional groups attached (as in glutamine).
The physical and chemical properties of the R group determine the unique
characteristics of a particular amino acid.
One group of amino acids has hydrophobic R groups.
Another group of amino acids has polar R groups that are hydrophilic.
A third group of amino acids includes those with functional groups that are charged
(ionized) at cellular pH.
Some acidic R groups are negative in charge due to the presence of a carboxyl group.
Basic R groups have amino groups that are positive in charge.
Note that all amino acids have carboxyl and amino groups. The terms acidic and basic
in this context refer only to these groups in the R groups.
Amino acids are joined together when a dehydration reaction removes a hydroxyl group
from the carboxyl end of one amino acid and a hydrogen from the amino group of
another.
The resulting covalent bond is called a peptide bond.
Repeating the process over and over creates a polypeptide chain.
At one end is an amino acid with a free amino group (the N-terminus) and at the other is
an amino acid with a free carboxyl group (the C-terminus).
Polypeptides range in size from a few monomers to thousands.
Each polypeptide has a unique linear sequence of amino acids.
He then searched for overlapping regions among the pieces obtained by hydrolyzing
with the different agents.
After years of effort, Sanger was able to reconstruct the complete primary structure of
insulin.
Most of the steps in sequencing a polypeptide have since been automated.
The mRNA molecule interacts with the cell’s protein-synthesizing machinery to direct
the ordering of amino acids in a polypeptide.
The flow of genetic information is from DNA -> RNA -> protein.
Protein synthesis occurs on cellular structures called ribosomes.
In eukaryotes, DNA is located in the nucleus, but most ribosomes are in the cytoplasm.
mRNA functions as an intermediary, moving information and directions from the nucleus
to the cytoplasm.
Prokaryotes lack nuclei but still use RNA as an intermediary to carry a message from
DNA to the ribosomes.
Module 3
Week 5
MOST ESSENTIAL LEARNING COMPETENCIES
1. Use simple collision theory to explain the effects of concentration, temperature,
and particle size on the rate of reaction.
2. Define catalyst and describe how its affects reaction rate.
3. Determine the limiting reactant in a reaction and calculate the amount of product
formed.
Rate of Reaction
Collision Theory which states that reactant particles require sufficient kinetic
energy to initiate successful collision that will lead to the formation of products.
The speed by which reactants are converted to products is referred to as the rate
of reaction. A reaction that takes a long time to complete is described to have a low
reaction rate. While some reactions may occur almost instantaneously, others may take
hours, days, or even one year to progress to completion.
Effect of Concentration
Reactant particles with the same phase (liquid-liquid or gas-gas) may easily mix
with each other. This gives the particles the maximum opportunity to collide and react.
The case is different when one of the reactants is a solid; the reaction can only take
place on the surface of the solid. The smaller the size of the solid particles, the greater
is the area in which the reaction can take place. Therefore, finely divided powder is
Physical Science
expected to react more quickly than a big limp of the same amount of the substance.
For example, wood shavings burn faster than big chunks of wood. Furthermore,
powdered metals react violently with strong acids compared to large pieces of the same
metal.
Effects of Temperature
When the reaction temperature goes up, the reaction rate increases as well. How
does this happen? The average kinetic energy of particles increases with rising
temperature, which means that at higher temperatures, particles tend to move faster.
Consequently, more effective collisions are favored. As depicted the fraction of the
reactant particles that can surpass the activation energy (represented by the shaded
portion of the graph) increases at temperature increases.
Effect of Catalyst
One of the important factors that affect the rate of a reaction is the presence of a
catalysts are substances that hasten reaction without themselves being consumed in
the reaction. They do so by lowering the activation energy that a reaction must
overcome in order to achieve successful collisions and progress.
In biological systems, catalysts are termed enzymes. They are proteins
considered absolutely essential to most life forms because many important chemical
reactions that occur in the body would progress much too slowly without them. Enzymes
are also used in some industrial processes such as the production of “biological
detergents”. This type of detergents may contain amylases to facilitate the
decomposition of starch, process for proteins, and lipases for fats and oils.
Physical Science
Module 4
Week 6
MOST ESSENTIAL LEARNING COMPETENCIES
1.Describe how energy is harnessed from different sources:
a. Fossil fuels
b. Biogas
c. Geothermal
d. Hydrothermal
e. Batteries
f. Solar cells
g. Biomass
Lecture:
Energy in Chemical Reaction
Due to the absorption of energy when chemical bonds are broken, and the release of
energy when chemical bonds are formed, chemical reactions almost always involve a
change in energy between products and reactants. By the Law of Conservation of
Energy, however, we know that the total energy of a system must remain unchanged,
Physical Science
and that oftentimes a chemical reaction will absorb or release energy in the form of
heat, light, or both. The energy change in a chemical reaction is due to the difference in
the amounts of stored chemical energy between the products and the reactants. This
stored chemical energy, or heat content, of the system is known as its enthalpy.
Exothermic Reactions
Exothermic reaction release heat and light into their surroundings. For example,
combustion reactions are usually exothermic. In exothermic reactions, the products
have less enthalpy than the reactants, and as a result, an exothermic reaction is said to
have a negative enthalpy of reaction. This means that the energy required to break the
bonds in the reactants is less than the energy released when new bonds form in the
products. Excess energy from the reaction is released as heat and light.
Endothermic Reactions
Endothermic reactions, on the other hand, absorb heat and/or light from their
surroundings. For example, decomposition reactions are usually endothermic. In
endothermic reactions, the products have more enthalpy than the reactants. Thus, an
endothermic reaction is said to have a positive enthalpy of reaction. This means that the
energy required to break the bonds in the reactants is more than the energy released
when new bonds form in the products; in other words, the reaction requires energy to
proceed.
Physical Science
1. Renewable Energy
a. Solar Energy
b. Wind Energy
c. Biomass and Biofuels
d. Water and Geothermal
2. Non Renewable
Energy is the capacity of a physical system to perform work. Energy exists in several
forms such as heat , kinetic or mechanical energy, light, potential energy , electrical, or
other forms. Energy is the ability to do work. Energy sources could be classified as
Renewable and Non-renewable.
Renewable Energy
Renewable energy is derived from natural processes that are replenished constantly
such as solar, wind, ocean, hydropower, biomass, geothermal resources, and biofuels
and hydrogen.
Solar Energy
Sun is the primary source of energy. Sunlight is a clean, renewable source of energy. It
is a sustainable resource, meaning it doesn't run out, but can be maintained because
the sun shines almost every day. Coal or gas are not sustainable or renewable: once
they are gone, there is none left. More and more people are wanting to use clean,
renewable energy such as solar, wind, geothermal steam and others. It is called 'Green
Power'. It lights our houses by day, dries our clothes and agricultural produce, keeps us
warm and lots more. Its potential is however much larger
Advantages
It is available in plenty
It is non-polluting
It does not emit any greenhouse gases.
Solar energy offers decentralization in most (sunny) locations, meaning self-
reliant societies.
One of the biggest advantages of solar energy is the ability to avoid the politics
and price volatility that is increasingly characterizing fossil fuel markets.
It doesn’t result in the destruction of forests and eco-systems that occurs with
most fossil fuel operations.
Disadvantages
It is environment friendly
Its freely and abundantly available
Disadvantages
Usage
Biomass is an important source of energy accounting for about one third of the total fuel
used in our country and in about 40% of the rural households. The widespread use of
biomass is for household cooking and heating. The types of biomass used are
agricultural waste, wood, charcoal or dried dung.
Advantages
Disadvantages
Water
The flowing water and the tides in the sea are sources of energy. India is endowed with
large hydropower potential of 1,45,320 MW. Heavy investments are made on large
projects. In recent years, hydel energy (through mini and small hydel power plants) is
Physical Science
also used to reach power to remote villages which are UN electrified. The estimated
potential of Small Hydro Power is about 15,000 MW in the country. As on May 2020, the
installed capacity of Small hydro projects (up to 3MW) amounts to 4683.16 MW
Advantages of Small Hydro Power as an energy source
Ocean energy
Oceans cover 70 percent of the earth’s surface and represent an enormous amount of
energy. Although currently under-utilized, Ocean energy is mostly exploited by just a
few technologies: Wave, Tidal, Current Energy and Ocean Thermal Energy.
1. Tidal Energy : The tidal cycle occurs every 12 hours due to the gravitational
force of the moon. The difference in water height from low tide and high tide is
potential energy. Similar to traditional hydropower generated from dams, tidal
water can be captured in a barrage across an estuary during high tide and forced
through a hydro-turbine during low tide. The capital cost for tidal energy power
plants is very high due to high civil construction and high power purchase tariff.
To capture sufficient power from the tidal energy potential, the height of high tide
must be at least five meters (16 feet) greater than low tide. Total identified
potential of Tidal Energy is about 12455 MW, with potential locations identified at
Khambat & Kutch regions, and large backwaters, where barrage technology
could be used.
2. Wave Energy : Wave energy is generated by the movement of a device either
floating on the surface of the ocean or moored to the ocean floor. Many different
techniques for converting wave energy to electric power have been studied.
Wave conversion devices that float on the surface have joints hinged together
that bend with the waves. This kinetic energy pumps fluid through turbines and
creates electric power. Stationary wave energy conversion devices use pressure
fluctuations produced in long tubes from the waves swelling up and down. This
bobbing motion drives a turbine when critical pressure is reached. Other
stationary platforms capture water from waves on their platforms. This water is
allowed to runoff through narrow pipes that flow through a typical hydraulic
Physical Science
turbine. The total theoretical potential of wave energy in India along the country’s
coast is estimated to be about 40,000 MW – these are preliminary estimates.
This energy is however less intensive than what is available in more northern and
southern latitudes.
3. Current Energy : Marine current is ocean water moving in one direction. This
ocean current is known as the Gulf Stream. Tides also create currents that flow
in two directions. Kinetic energy can be captured from the Gulf Stream and other
tidal currents with submerged turbines that are very similar in appearance to
miniature wind turbines. Similar to wind turbines, the movement of the marine
current moves the rotor blades to generate electric power.
4. Ocean Thermal Energy Conversion (OTEC): Ocean thermal energy
conversion, or OTEC, uses ocean temperature differences from the surface to
depths lower than 1,000 meters, to extract energy. A temperature difference of
only 20°C can yield usable energy. Research focuses on two types of OTEC
technologies to extract thermal energy and convert it to electric power: closed
cycle and open cycle. In the closed cycle method, a working fluid, such as
ammonia, is pumped through a heat exchanger and vaporized. This vaporized
steam runs a turbine. The cold water found at the depths of the ocean condenses
the vapor back to a fluid where it returns to the heat exchanger. In the open cycle
system, the warm surface water is pressurized in a vacuum chamber and
converted to steam to run the turbine. The steam is then condensed using cold
ocean water from lower depths. OTEC has a theoretical potential of 180,000 MW
in India subject to suitable technological evolution.
Geothermal energy
Geothermal Energy is heat stored in earth crust and being used for electric generation
and also for direct heat application. Geothermal literally means heat generated by earth.
Various resource assessment carried out by agencies established the potential 10600
MWth /1000MWe spread over 340 hot springs across seven Geothermal provinces/11
states.
The availability of geothermal power is most environment-friendly power, round the year
24x7 basis, not affected by the severity of climate during 6 to 7 winter months like hydro
and like dependence on sun in solar PV.
.
Non Renewable energy
Coal, Oil and Natural gas are the non-renewable sources of energy. They are also
called fossil fuels as they are products of plants that lived thousands of years ago.
Fossil fuels are the predominantly used energy sources today. India is the third largest
producer of coal in the world, with estimated reserves of around 31,902,033 million
tunes of Geological Resources of Coal (as of 1.4.2018). Coal supplies more than
70.87% of the country's total production of energy by commercial sources. India
consumes about 245 MT of crude oil annually, and more than 70% of it is imported.
Burning fossil fuels cause great amount of environmental pollution.
Physical Science
Module 5
Week 7 and 8
MOST ESSENTIAL LEARNING COMPETENCIES
1. From products labels, identify the active ingredient(s) of cleaning products used
at home.
2. Give the use of the other ingredients in cleaning agents.
Lecture
Consumer Products
The most common consumer products are intended for household cleaning or
personal care. These products contain substances that may be classified as active or
inactive ingredients. Active ingredients are substances that directly help in achieving
the performance objectives of a certain product. Other components though, which
usually constitute a large percentage of the product, may be considered as inactive or
inert ingredients like fragrance and solvents.
The main function of consumer products depends on the ability of the active
ingredients to react with target substances. For example, if a product is claimed to
remove stains in fabric, it should contains ingredients that actively react with stain and
make its removal from cloth easier.
Household Cleaning Products
Different household cleaners contain substances that may fall under specific
types, including surfactants, bleaching agent and disinfectants. These substances may
undergo chemical reactions that render the product effective.
Type and Ingredients of Household Cleaning Products
Household cleaning products are generally classified according to their purpose
or specific application, which includes toilet and bathroom cleaning, stain removal, and
sanitation. This manner of classifying is somehow arbitrary because many products
offer a variety uses. Some types of cleaning products, including their common
ingredients, are briefly described in the succeeding discussion. Knowing these
ingredients is helpful, especially when comparing product performance and safety.
Bathroom Cleaners
Bathroom cleaner are products specifically designed for bathroom
surfaces, such as tubs, tiles, and toilet bowls, which normally develop
stains like mildew. Bathroom cleaners are either acidic or alkaline
depending on the surface they will be used on. Alkaline cleaners are best
to use for bathroom floors, walls, tiles, and bathtubs because they
preserve the enamel finishes of these surfaces; acidic cleanser can
damage enamels. Cleansers with acidic active ingredients like phosphoric
acid and citric acid are recommended for other bathroom parts with rust
and mineral deposits.
Special Surface Cleanser
Physical Science
Different cleaning agents are used depending on the item to be cleaned, the cleaning
method and the type of soiling found on the item. There are four main types of cleaning
agents used in commercial kitchens:
1. Detergents
2. Degreasers
3. Abrasives
4. Acids
Physical Science
Detergents
Detergents are the most common type of cleaning agent and are used in home and
commercial kitchens. They work by breaking up dirt or soil, making it easy to wash it
away.
The detergents used in commercial kitchens are usually synthetic detergents made from
petroleum products and may be in the form of powder, liquid, gel or crystals.
Degreasers
Degreasers are sometimes known as solvent cleaners and are used to remove grease
from surfaces such as oven tops, counters and grill backsplashes.
Methylated spirits or white spirit were commonly used as degreasers in the past. Most
food businesses now try to use non-toxic, non-fuming degreasers in their operations to
prevent chemical contamination.
Abrasives
Abrasives should be used with care as they may scratch certain types of materials used
for kitchen equipment such as plastic or stainless steel.
Acids
Acid cleaners are the most powerful type of cleaning agent and should be used with
care. If they are not diluted correctly acid cleaners can be very poisonous and corrosive.
Acid cleaners are generally used to remove mineral deposits and are useful for
descaling dishwashers or removing rust from restroom facilities.
Cleaning is only the first step to a germ-free kitchen. Cleaning is done using detergent,
but it doesn’t kill bacteria or other microorganisms that can cause food poisoning. To kill
bacteria and ensure a clean workplace, you must follow cleaning with sanitizing.
prevent cross-contamination
hazards. If they are used, stored and disposed of properly, however, they can be
relatively safe.
Things to make your home safer
Recycle
Many products are recyclable. Contact your recycling coordinator or local department of
environmental services to find out what is being recycled in your community.
Dispose of properly
Products should NEVER be discarded on the ground or poured into storm drains. Many
products shouldn't even be disposed of in the trash or down the toilet. These products
should be saved and taken to Household Hazardous Waste (HHW) collections. Contact
your sanitation department, local or state department of environmental services for
information on HHW collections in your area.
Module 6
Week 9 and 10
MOST ESSENTIAL LEARNING COMPETENCIES
1 Explain how the Greeks knew that the Earth is Spherical.
2 Cite examples of astronomical phenomena known to astronomical before the advent
of telescopes
3 Explain how Brahe’s innovation and extensive collection of data in observational
astronomy paved the way of Kepler’s discovery of his laws of planetary motion.
Lecture:
The spherical Earth
Greeks knew the Earth was a sphere
View of constellations changes from North to South
Observations of ships sailing over the horizon (mast disappears last)
Observations of the Earth’s shadow on the Moon during lunar eclipses
The myth of the Flat Earth is a modern misconception about ancient and
medieval thought
This is a bit
oversimplified!
Note mountains not
shown to scale.
Aether
Aristotle’s celestial physics
Heavens are governed by different laws from Earth
Celestial bodies composed of “aether,” a fifth element not present on Earth*
*turns out there might a quintessence! (later…)
“Natural motions” of celestial spheres are different from terrestrial motions:
circular, constant, and eternal
Physical Science
Epic Epicycles
Aristotelian/Ptolemaic view prevailed in Europe and Islamic empire, through
1400’s
Geocentric model
Creation at finite time in past, for consistency with Christian theology
Earth known to be round (Columbus battling against flat Earthers is myth!)
Copernicus (1473-1543)
Nicholas Copernicus was modern (re)founder of the heliocentric (Sun centered)
model for the solar system
Rejected Ptolemy’s geocentric model because it was too complicated (Occam's
razor)
Heliocentric model …
with perfect circular motions (model bias!)
Mathematics was not simpler than Ptolemy's, but it required fewer basic
assumptions
By Copernicus assuming only
the rotation of the Earth,
revolution about the sun,
tilt of Earth's rotational axis,
he could explain observed “motions” of the heavens
Because he retained circular orbits, his system required the inclusion of epicycles
The Copernican Principle: The Earth is not at a special location in the Universe.
Later, we will come across the Generalized Copernican Principle: There is no
special place in the universe, i.e., the universe has no center
This was far more than the income of any other man of learning in Europe, and
even for an aristocrat, it was a substantial income.
Tycho set a new European standard for the financial support of scientific
research.
It is estimated that Brahe's observatory cost about 1% of the Danish government
budget during construction.
Tycho’s Compromise
Alas, stellar parallax is 100× too small for naked-eye observation to
measure; largest values are about 1”
1 arcsecond=(1/3600)°
What’s 1” in radians? Board digression
How many AU away is 1 arcsecond?
What is the parallax of α-Centauri, 4 ly away?
Kepler in perspective
•Based on Tycho Brahe’s accurate observations, Kepler calculated his way to a
major breakthrough in cosmology
Kepler’s three laws of planetary motion
Represented a simpler model of solar system and can be generalized
Swept away thousands of years of prejudice – and his own previous pet
theory!
Were driven fundamentally by the data, including Tycho’s error estimates
Data will lead theory again in cosmology!
Unlike previous models, Kepler’s Laws had predictive power, consistent with
modern idea of a meaningful scientific theory
Newton’s derivation (i.e., deeper underlying theory) of Kepler’s laws from even
more general principles comes later…
19th century observations of small deviation of Mercury’s orbit from
Newton’s/Kepler's laws spur theoreticians again… more on that later
Module 7
Week 11
1 Compare and contrast the Aristotelian and Gallilean conceptions of vertical motion,
horizontal motion, and projectile motion.
2 Explain how Galileo inferred that objects in vacuum fall uniform acceleration, and that
force is not necessary to sustain horizontal motion.
3. Explain the subtle distinction between Newton’s 1st Law of motion ( or Law of Inertia )
and Gallileo’s assertion that firce is not necessary to sustain horizontal motion.
Physical Science
Who is Aristotle?
Aristotle was an ancient Greek philosopher and scientist born in the city of Stagira,
Chalkidice, on the northern periphery of Classical Greece.
Aristotle was considered the most outstanding philosopher-scientist of his time in
ancient Greece.
Aristotle’s Ideas on Motion
» He defines motion as the actuality of a potentiality. Initially, Aristotle’s definition seems
to involve a contradiction. However, commentators on the works of Aristotle, such as St.
Thomas Aquinas, maintain that this is the only way to define motion.
» Aristotle divided motion into two main classes: natural motion and violent motion
Natural Motion:
♥ Any motion that an object does naturally – without being forced – was classified by
Aristotle as a natural motion.
♥ Solid objects (or liquids) fall because they seek their natural resting place (centerof
the Earth).
♥ Air likes to rise upwards, as do flames, since that is their natural resting place.
The sun naturally rises in the east, crosses the sky, then sets in the west.
Violent Motion:
Aristotle classified any motion that required a force as a “violent motion”. (He did not
mean violent in the modern sense…)
♠ Horizontal motion
♠ Imposed by pushes or pulls; objects move not by their nature, but because of
impressed forces
.
Aristotle’s views on Motion
Aristotle’s observations
VERTICAL MOTION
– The element earth moves down toward its natural resting place.
– Water’s natural place is just above earth.
– Air rises to its natural place in the atmosphere.
– Fire leaps upwards to its natural place above the atmosphere.
HORIZONTAL MOTION
– Qualitatively different.
– Bodies seem to need push or pull to maintain horizontal motion
(contrary to their ‘natural’ motion).
PROJECTILE MOTION
– is an object upon which the only force acting is gravity.
– A projectile is any object that once projected or dropped continues in motion by its
own inertia and is influenced only by the downward force of gravity.
Physical Science
He also had his view on the projectile motion of an object. He believed that an object
thrown at a certain angle is given an impetus—a force or energy that permits an object
to move. It will continue to move in such state until the object’s impetus is lost, and the
object returns to its natural state, causing it to stop and fall to the ground.
Galileo disproved Aristotle’s claims and believed that the motion of objects is not simply
due to the composition of objects. He mentioned that motion can be described by
mathematics and the changes in some physical variables such as time and distance.
Using his actual and thorough experiments, he was able to prove that:
1. an object in uniform motion will travel a distance that is proportional to the time it will
take to travel;
2. a uniformly accelerating object will travel at a speed proportional to some factor of time;
and
3. an object in motion, if unimpeded, will continue to be in motion; an external force is not
necessary to maintain the motion.
Horizontal motion
Vertical motion
In the absence of a resistance, objects would fall not depending on their weight, but in
the time of fall. Also, if the object encountered a resistive force from a fluid equal or
greater than its weight, it will slow down and reaches a uniform motion until it reaches
the bottom and stops. For example, without any resistance, a 1-kg object will be as fast
as a 10-kg object when falling because they fall with the same amount of time, given
that they are released from the same height. Also, a stone dropped in the ocean will
sooner or later travel at constant speed.
Projectile motion
Try it!
Who has a more acceptable view of falling objects, Aristotle or Galileo? Try to test it in
the following activities below. Take note that in every activity, both objects should be
dropped at the same time and at the same height.
Assesment I
A. I and III
B. II and IV
C. II and III
D. All of the above
5. Which of the following agrees with Galileo’s view of motion?
A. A feather will be as fast as an iron ball if dropped in a vacuum.
B. If a 10-kg boulder is dropped to the sea, it will stop midway.
C. Planets revolve around the Sun because of a Prime Mover.
D. Smoke rises because the sky is its natural place.
6. Which of the following is true about the projectile of an arrow when shot?
A. It has a continuous applied force from the person who shot the arrow.
B. It creates a vacuum that sucks air in, and the air pushes the arrow.
C. It rises because arrows are mostly air.
D. It has both uniform motion and uniformly accelerating motion.
7. Which of the following agrees with Galileo’s view of motion?
A. A flying baseball has both uniform motion and uniformly accelerating motion.
B. A box will stop sliding in a frictionless plane if the applied force is removed.
C. A balloon rises because it is mostly air, and the sky is its natural place.
D. An iron ball will always fall faster than a rubber ball.
8. According to Galileo's view, which will reach the floor of the vacuum chamber first, a
sheet of paper or a bowling ball?
A. The bowling ball will reach the vacuum chamber.
B. The sheet of paper will reach the vacuum chamber.
C. They will reach the floor at the same time.
D. They will not reach the floor and will be sucked in the vacuum.
A box sliding on the floor slowly stops.
9. Why is this so?
(i) Mass,
(ii) Motion &
(iii) Force
Newton's First Law
First Law
Inertia and Mass
State of Motion
Balanced vs. Unbalanced Forces
There are two clauses or parts to this statement - one that predicts the behavior of
stationary objects and the other that predicts the behavior of moving objects. The two
parts are summarized in the following diagram.
Newton's second law of motion pertains to the behavior of objects for which all existing
forces are not balanced. The second law states that the acceleration of an object is
dependent upon two variables - the net force acting upon the object and the mass of the
object. The acceleration of an object depends directly upon the net force acting upon
the object, and inversely upon the mass of the object. As the force acting upon an
object is increased, the acceleration of the object is increased. As the mass of an object
is increased, the acceleration of the object is decreased.
The above equation is often rearranged to a more familiar form as shown below. The
net force is equated to the product of the mass times the acceleration.
Fnet = m • a
In this entire discussion, the emphasis has been on the net force. The
acceleration is directly proportional to the net force; the net force equals
mass times acceleration; the acceleration in the same direction as
the net force; an acceleration is produced by a net force. The NET
FORCE. It is important to remember this distinction. Do not use the
value of merely "any 'ole force" in the above equation. It is the net force
that is related to acceleration. As discussed in an earlier lesson, the net
force is the vector sum of all the forces. If all the individual forces acting
upon an object are known, then the net force can be determined. If necessary, review
this principle by returning to the practice questions in Lesson 2.
Consistent with the above equation, a unit of force is equal to a unit of mass times a unit
of acceleration. By substituting standard metric units for force, mass, and acceleration
into the above equation, the following unit equivalency can be written.
1 Newton = 1 kg • m/s2
The definition of the standard metric unit of force is stated by the above equation. One
Newton is defined as the amount of force required to give a 1-kg mass an acceleration
of 1 m/s/s.