PROPERTIES OF SOLIDS AND LIQUIDS – 3

Cohesion and Adhesion in Liquids: Surface Tension and Capillary Action

Children blow soap bubbles and play in the spray of a sprinkler on a hot summer day. (See Figure below ) . A technician draws blood into a small-diameter tube just by touching it to a drop on a pricked finger. A premature infant struggles to inflate her lungs. What is the common thread? All these activities are dominated by the attractive forces between atoms and molecules in liquids—both within a liquid and between the liquid and its surroundings.

Attractive forces between molecules of the same type are called cohesive forces. Liquids can, for example, be held in open containers because cohesive forces hold the molecules together. Attractive forces between molecules of different types are called adhesive forces. Such forces cause liquid drops to cling to window panes, for example. Lets examine effects directly attributable to cohesive and adhesive forces in liquids.

Cohesive Forces
Attractive forces between molecules of the same type are called cohesive forces.

Adhesive Forces
Attractive forces between molecules of different types are called adhesive forces.

The soap bubbles in this photograph are caused by cohesive forces among molecules in liquids. (credit: Steve Ford Elliott)

Surface Tension

Cohesive forces between molecules cause the surface of a liquid to contract to the smallest possible surface area. This general effect is called surface tension. Molecules on the surface are pulled inward by cohesive forces, reducing the surface area. Molecules inside the liquid experience zero net force, since they have neighbors on all sides.

Surface Tension
Cohesive forces between molecules cause the surface of a liquid to contract to the smallest possible surface area. This general effect is called surface tension.

The model of a liquid surface acting like a stretched elastic sheet can effectively explain surface tension effects. For example, some insects can walk on water (as opposed to floating in it) as we would walk on a trampoline—they dent the surface as shown in the below figure (a). The next figure (b) shows another example, where a needle rests on a water surface. The iron needle cannot, and does not, float, because its density is greater than that of water. Rather, its weight is supported by forces in the stretched surface that try to make the surface smaller or flatter. If the needle were placed point down on the surface, its weight acting on a smaller area would break the surface, and it would sink.

Surface tension supporting the weight of an insect and an iron needle, both of which rest on the surface without penetrating it. They are not floating; rather, they are supported by the surface of the liquid. (a) An insect leg dents the water surface.  is a restoring force (surface tension) parallel to the surface. (b) An iron needle similarly dents a water surface until the restoring force (surface tension) grows to equal its weight.

Surface tension is proportional to the strength of the cohesive force, which varies with the type of liquid. Surface tension  is defined to be the force F per unit length  exerted by a stretched liquid membrane

The tables below lists values of  for some liquids.

For the insect of the above figure (a), its weight w is supported by the upward components of the surface tension force: w=γLsinθ, where L is the circumference of the insect’s foot in contact with the water. The figure below shows one way to measure surface tension. The liquid film exerts a force on the movable wire in an attempt to reduce its surface area. The magnitude of this force depends on the surface tension of the liquid and can be measured accurately.

Sliding wire device used for measuring surface tension; the device exerts a force to reduce the film’s surface area. The force needed to hold the wire in place is F=γL=γ(2l), since there are two liquid surfaces attached to the wire. This force remains nearly constant as the film is stretched, until the film approaches its breaking point

Surface tension is the reason why liquids form bubbles and droplets. The inward surface tension force causes bubbles to be approximately spherical and raises the pressure of the gas trapped inside relative to atmospheric pressure outside. It can be shown that the gauge pressure P inside a spherical bubble is given by

where r is the radius of the bubble. Thus the pressure inside a bubble is greatest when the bubble is the smallest. Another bit of evidence for this is illustrated in the figure below . When air is allowed to flow between two balloons of unequal size, the smaller balloon tends to collapse, filling the larger balloon

With the valve closed, two balloons of different sizes are attached to each end of a tube. Upon opening the valve, the smaller balloon decreases in size with the air moving to fill the larger balloon. The pressure in a spherical balloon is inversely proportional to its radius, so that the smaller balloon has a greater internal pressure than the larger balloon, resulting in this flow

Our lungs contain hundreds of millions of mucus-lined sacs called alveoli, which are very similar in size, and about 0.1 mm in diameter. (See figure below.) You can exhale without muscle action by allowing surface tension to contract these sacs. Medical patients whose breathing is aided by a positive pressure respirator have air blown into the lungs, but are generally allowed to exhale on their own. Even if there is paralysis, surface tension in the alveoli will expel air from the lungs. Since pressure increases as the radii of the alveoli decrease, an occasional deep cleansing breath is needed to fully reinflate the alveoli. Respirators are programmed to do this and we find it natural

Bronchial tubes in the lungs branch into ever-smaller structures, finally ending in alveoli. The alveoli act like tiny bubbles. The surface tension of their mucous lining aids in exhalation and can prevent inhalation if too great

The tension in the walls of the alveoli results from the membrane tissue and a liquid on the walls of the alveoli containing a long lipoprotein that acts as a surfactant (a surface-tension reducing substance). The need for the surfactant results from the tendency of small alveoli to collapse and the air to fill into the larger alveoli making them even larger . During inhalation, the lipoprotein molecules are pulled apart and the wall tension increases as the radius increases (increased surface tension). During exhalation, the molecules slide back together and the surface tension decreases, helping to prevent a collapse of the alveoli. The surfactant therefore serves to change the wall tension so that small alveoli don’t collapse and large alveoli are prevented from expanding too much. This tension change is a unique property of these surfactants, and is not shared by detergents (which simply lower surface tension). (See figure below)

Surface tension as a function of surface area. The surface tension for lung surfactant decreases with decreasing area. This ensures that small alveoli don’t collapse and large alveoli are not able to over expand

If water gets into the lungs, the surface tension is too great and you cannot inhale. This is a severe problem in resuscitating drowning victims. A similar problem occurs in newborn infants who are born without this surfactant—their lungs are very difficult to inflate. This condition is known as hyaline membrane disease and is a leading cause of death for infants, particularly in premature births. Some success has been achieved in treating hyaline membrane disease by spraying a surfactant into the infant’s breathing passages. Emphysema produces the opposite problem with alveoli. Alveolar walls of emphysema victims deteriorate, and the sacs combine to form larger sacs. Because pressure produced by surface tension decreases with increasing radius, these larger sacs produce smaller pressure, reducing the ability of emphysema victims to exhale. A common test for emphysema is to measure the pressure and volume of air that can be exhaled.

Adhesion and Capillary Action

Why is it that water beads up on a waxed car but does not on bare paint? The answer is that the adhesive forces between water and wax are much smaller than those between water and paint. Competition between the forces of adhesion and cohesion are important in the macroscopic behavior of liquids. An important factor in studying the roles of these two forces is the angle θ between the tangent to the liquid surface and the surface. (See the figure below .) The contact angle  θ is directly related to the relative strength of the cohesive and adhesive forces. The larger the strength of the cohesive force relative to the adhesive force, the larger θ is, and the more the liquid tends to form a droplet. The smaller θ is, the smaller the relative strength, so that the adhesive force is able to flatten the drop

Contact Angle
The angle θ between the tangent to the liquid surface and the surface is called the contact angle.

In the photograph, water beads on the waxed car paint and flattens on the unwaxed paint. (a) Water forms beads on the waxed surface because the cohesive forces responsible for surface tension are larger than the adhesive forces, which tend to flatten the drop. (b) Water beads on bare paint are flattened considerably because the adhesive forces between water and paint are strong, overcoming surface tension. The contact angle θ is directly related to the relative strengths of the cohesive and adhesive forces. The larger θ is, the larger the ratio of cohesive to adhesive forces. (credit: P. P. Urone)

.  The table below lists contact angles for several combinations of liquids and solids.

One important phenomenon related to the relative strength of cohesive and adhesive forces is capillary action—the tendency of a fluid to be raised or suppressed in a narrow tube, or capillary tube. This action causes blood to be drawn into a small-diameter tube when the tube touches a drop.

Capillary Action
The tendency of a fluid to be raised or suppressed in a narrow tube, or capillary tube, is called capillary action.

If a capillary tube is placed vertically into a liquid, as shown in the figure below , capillary action will raise or suppress the liquid inside the tube depending on the combination of substances. The actual effect depends on the relative strength of the cohesive and adhesive forces and, thus, the contact angle θ given in the table. If θ is less than 900,  then the fluid will be raised; if θ is greater than 900, it will be suppressed. Mercury, for example, has a very large surface tension and a large contact angle with glass. When placed in a tube, the surface of a column of mercury curves downward, somewhat like a drop. The curved surface of a fluid in a tube is called a meniscus. The tendency of surface tension is always to reduce the surface area. Surface tension thus flattens the curved liquid surface in a capillary tube. This results in a downward force in mercury and an upward force in water, as seen in the figure below .

(a) Mercury is suppressed in a glass tube because its contact angle is greater than . Surface tension exerts a downward force as it flattens the mercury, suppressing it in the tube. The dashed line shows the shape the mercury surface would have without the flattening effect of surface tension. (b) Water is raised in a glass tube because its contact angle is nearly . Surface tension therefore exerts an upward force when it flattens the surface to reduce its area.

Capillary action can move liquids horizontally over very large distances, but the height to which it can raise or suppress a liquid in a tube is limited by its weight. It can be shown that this height  is given by

If we look at the different factors in this expression, we might see how it makes good sense. The height is directly proportional to the surface tension , which is its direct cause. Furthermore, the height is inversely proportional to tube radius—the smaller the radius , the higher the fluid can be raised, since a smaller tube holds less mass. The height is also inversely proportional to fluid density , since a larger density means a greater mass in the same volume. (See figure below.)

(a) Capillary action depends on the radius of a tube. The smaller the tube, the greater the height reached. The height is negligible for large-radius tubes. (b) A denser fluid in the same tube rises to a smaller height, all other factors being the same

Thermal Expansion of Solids and Liquids

The expansion of alcohol in a thermometer is one of many commonly encountered examples of thermal expansion, the change in size or volume of a given mass with temperature. Hot air rises because its volume increases, which causes the hot air’s density to be smaller than the density of surrounding air, causing a buoyant (upward) force on the hot air. The same happens in all liquids and gases, driving natural heat transfer upwards in homes, oceans, and weather systems. Solids also undergo thermal expansion. Railroad tracks and bridges, for example, have expansion joints to allow them to freely expand and contract with temperature changes.

Thermal expansion joints like these in the Auckland Harbour Bridge in New Zealand allow bridges to change length without buckling. (credit: Ingolfson, Wikimedia Commons)

First, thermal expansion is clearly related to temperature change. The greater the temperature change, the more a bimetallic strip will bend. Second, it depends on the material. In a thermometer, for example, the expansion of alcohol is much greater than the expansion of the glass containing it.

an increase in temperature implies an increase in the kinetic energy of the individual atoms. In a solid, unlike in a gas, the atoms or molecules are closely packed together, but their kinetic energy (in the form of small, rapid vibrations) pushes neighboring atoms or molecules apart from each other. This neighbor-to-neighbor pushing results in a slightly greater distance, on average, between neighbors, and adds up to a larger size for the whole body. For most substances under ordinary conditions, there is no preferred direction, and an increase in temperature will increase the solid’s size by a certain fraction in each dimension

The change in length ∆ L is proportional to length L. The dependence of thermal expansion on temperature, substance, and length is summarized in the equation

∆ L = α L ∆ T

Where ∆ L  is the change in length , ∆ T  is the change in temperature, and α  is the coefficient of linear expansion, which varies slightly with temperature.

The below table  lists representative values of the coefficient of linear expansion, which may have units of 1/0 C or 1/K. Because the size of a kelvin and a degree Celsius are the same, both and can be expressed in units of kelvins or degrees Celsius. The equation ∆ L = α L ∆ T  is accurate for small changes in temperature and can be used for large changes in temperature if an average value of is used.

Thermal Expansion in Two and Three Dimensions

Objects expand in all dimensions, as illustrated in the below figure . That is, their areas and volumes, as well as their lengths, increase with temperature. Holes also get larger with temperature. If you cut a hole in a metal plate, the remaining material will expand exactly as it would if the plug was still in place. The plug would get bigger, and so the hole must get bigger too. (Think of the ring of neighboring atoms or molecules on the wall of the hole as pushing each other farther apart as temperature increases. Obviously, the ring of neighbors must get slightly larger, so the hole gets slightly larger).

For small temperature changes, the change in area ∆ A is given by

∆ A = 2 α A ∆ T

Where ∆ A is the change in area , ∆ T is the change in temperature, and α is the coefficient of linear expansion, which varies slightly with temperature

In general, objects expand in all directions as temperature increases. In these drawings, the original boundaries of the objects are shown with solid lines, and the expanded boundaries with dashed lines. (a) Area increases because both length and width increase. The area of a circular plug also increases. (b) If the plug is removed, the hole it leaves becomes larger with increasing temperature, just as if the expanding plug were still in place. (c) Volume also increases, because all three dimensions increase

In general, objects will expand with increasing temperature. Water is the most important exception to this rule. Water expands with increasing temperature (its density decreases) when it is at temperatures greater than 4oC (40oF). However, it expands with decreasing temperature when it is between +4oC and 0oC(40oF 32oF). Water is densest at +4oC. (See below graph.) Perhaps the most striking effect of this phenomenon is the freezing of water in a pond. When water near the surface cools down to 4oC  it is denser than the remaining water and thus will sink to the bottom. This “turnover” results in a layer of warmer water near the surface, which is then cooled.

Eventually the pond has a uniform temperature of 4oC. If the temperature in the surface layer drops below 4oC, the water is less dense than the water below, and thus stays near the top. As a result, the pond surface can completely freeze over. The ice on top of liquid water provides an insulating layer from winter’s harsh exterior air temperatures. Fish and other aquatic life can survive in 4oC water beneath ice, due to this unusual characteristic of water. It also produces circulation of water in the pond that is necessary for a healthy ecosystem of the body of water.

The density of water as a function of temperature. Note that the thermal expansion is actually very small. The maximum density at +4oC is only 0.0075% greater than the density at 2oC, and 0.012% greater than that at 0oC.

Thermal Stress

Thermal stress is created by thermal expansion or contraction . Thermal stress can be destructive, such as when expanding petrol ruptures a tank. It can also be useful, for example, when two parts are joined together by heating one in manufacturing, then slipping it over the other and allowing the combination to cool. Thermal stress can explain many phenomena, such as the weathering of rocks and pavement by the expansion of ice when it freezes.

Forces and pressures created by thermal stress are typically large (See below figure.)

Thermal stress contributes to the formation of potholes. credit: Editor5807, Wikimedia Commons

Power lines sag more in the summer than in the winter, and will snap in cold weather if there is insufficient slack. Cracks open and close in plaster walls as a house warms and cools. Glass cooking pans will crack if cooled rapidly or unevenly, because of differential contraction and the stresses it creates. (Pyrex® is less susceptible because of its small coefficient of thermal expansion.) Nuclear reactor pressure vessels are threatened by overly rapid cooling, and although none have failed, several have been cooled faster than considered desirable. Biological cells are ruptured when foods are frozen, detracting from their taste. Repeated thawing and freezing accentuate the damage. Even the oceans can be affected. A significant portion of the rise in sea level that is resulting from global warming is due to the thermal expansion of sea water.

Metal is regularly used in the human body for hip and knee implants. Most implants need to be replaced over time because, among other things, metal does not bond with bone. Researchers are trying to find better metal coatings that would allow metal-to-bone bonding. One challenge is to find a coating that has an expansion coefficient similar to that of metal. If the expansion coefficients are too different, the thermal stresses during the manufacturing process lead to cracks at the coating-metal interface.

Another example of thermal stress is found in the mouth. Dental fillings can expand differently from tooth enamel. It can give pain when eating ice cream or having a hot drink. Cracks might occur in the filling. Metal fillings (gold, silver, etc.) are being replaced by composite fillings (porcelain), which have smaller coefficients of expansion, and are closer to those of teeth

Heat

 Work is defined as force times distance and in earlier chapters we learned that work done on an object changes its kinetic energy. We also saw in previous lessons that temperature is proportional to the (average) kinetic energy of atoms and molecules. We say that a thermal system has a certain internal energy: its internal energy is higher if the temperature is higher. If two objects at different temperatures are brought in contact with each other, energy is transferred from the hotter to the colder object until equilibrium is reached and the bodies reach thermal equilibrium (i.e., they are at the same temperature). No work is done by either object, because no force acts through a distance. The transfer of energy is caused by the temperature difference, and ceases once the temperatures are equal. These observations lead to the following definition of heat: Heat is the spontaneous transfer of energy due to a temperature difference

Heat is often confused with temperature. For example, we may say the heat was unbearable, when we actually mean that the temperature was high. Heat is a form of energy, whereas temperature is not. The misconception arises because we are sensitive to the flow of heat, rather than the temperature.

Owing to the fact that heat is a form of energy, it has the SI unit of joule (J). The calorie (cal) is a common unit of energy, defined as the energy needed to change the temperature of 1.00 g of water by 1.00oC —specifically, between 14.5oC and 15.5oC , since there is a slight temperature dependence. Perhaps the most common unit of heat is the kilocalorie (kcal), which is the energy needed to change the temperature of 1.00 kg of water by 1.00oC . Since mass is most often specified in kilograms, kilocalorie is commonly used. Food calories (given the notation Cal, and sometimes called “big calorie”) are actually kilocalories (1 Kilocalorie = 1000 calories), a fact not easily determined from package labeling.

Mechanical Equivalent of Heat

It is also possible to change the temperature of a substance by doing work. Work can transfer energy into or out of a system. This realization helped establish the fact that heat is a form of energy. James Prescott Joule (1818–1889) performed many experiments to establish the mechanical equivalent of heatthe work needed to produce the same effects as heat transfer. In terms of the units used for these two terms, the best modern value for this equivalence is

We consider this equation as the conversion between two different units of energy

Schematic depiction of Joule’s experiment that established the equivalence of heat and work

The figure above shows one of Joule’s most famous experimental setups for demonstrating the mechanical equivalent of heat. It demonstrated that work and heat can produce the same effects, and helped establish the principle of conservation of energy. Gravitational potential energy (PE) (work done by the gravitational force) is converted into kinetic energy (KE), and then randomized by viscosity and turbulence into increased average kinetic energy of atoms and molecules in the system, producing a temperature increase. His contributions to the field of thermodynamics were so significant that the SI unit of energy was named after him.

Heat added or removed from a system changes its internal energy and thus its temperature. Such a temperature increase is observed while cooking. However, adding heat does not necessarily increase the temperature. An example is melting of ice; that is, when a substance changes from one phase to another. Work done on the system or by the system can also change the internal energy of the system. Joule demonstrated that the temperature of a system can be increased by stirring. If an ice cube is rubbed against a rough surface, work is done by the frictional force. A system has a well-defined internal energy, but we cannot say that it has a certain “heat content” or “work content”. We use the phrase “heat transfer” to emphasize its nature.

Temperature Change and Heat Capacity

One of the major effects of heat transfer is temperature change: heating increases the temperature while cooling decreases it. We assume that there is no phase change and that no work is done on or by the system. Experiments show that the transferred heat depends on three factors—the change in temperature, the mass of the system, and the substance and phase of the substance.

 

The heat Q transferred to cause a temperature change depends on the magnitude of the temperature change, the mass of the system, and the substance and phase involved. (a) The amount of heat transferred is directly proportional to the temperature change. To double the temperature change of a mass m, you need to add twice the heat. (b) The amount of heat transferred is also directly proportional to the mass. To cause an equivalent temperature change in a doubled mass, you need to add twice the heat. (c) The amount of heat transferred depends on the substance and its phase. If it takes an amount Q of heat to cause a temperature change ∆T in a given mass of copper, it will take 10.8 times that amount of heat to cause the equivalent temperature change in the same mass of water assuming no phase change in either substance

The dependence on temperature change and mass are easily understood. Owing to the fact that the (average) kinetic energy of an atom or molecule is proportional to the absolute temperature, the internal energy of a system is proportional to the absolute temperature and the number of atoms or molecules. Owing to the fact that the transferred heat is equal to the change in the internal energy, the heat is proportional to the mass of the substance and the temperature change. The transferred heat also depends on the substance so that, for example, the heat necessary to raise the temperature is less for alcohol than for water. For the same substance, the transferred heat also depends on the phase (gas, liquid, or solid).

Values of specific heat must generally be looked up in tables, because there is no simple way to calculate them. In general, the specific heat also depends on the temperature. The table below lists representative values of specific heat for various substances. Except for gases, the temperature and volume dependence of the specific heat of most substances is weak. We see from this table that the specific heat of water is five times that of glass and ten times that of iron, which means that it takes five times as much heat to raise the temperature of water the same amount as for glass and ten times as much heat to raise the temperature of water as for iron. In fact, water has one of the largest specific heats of any material, which is important for sustaining life on Earth.

Calorimetry

One technique we can use to measure the amount of heat involved in a physical or chemical process is known as calorimetry. Calorimetry is used to measure amounts of heat transferred to or from a substance. To do so, the heat is exchanged with a calibrated object (calorimeter). The temperature change measured by the calorimeter is used to derive the amount of heat transferred by the process under study. The measurement of heat transfer using this approach requires the definition of a system (the substance or substances undergoing the chemical or physical change) and its surroundings (all other matter, including components of the measurement apparatus, that serve to either provide heat to the system or absorb heat from the system).

A calorimeter is a device used to measure the amount of heat involved in a chemical or physical process. For example, when an exothermic reaction occurs in solution in a calorimeter, the heat produced by the reaction is absorbed by the solution, which increases its temperature. When an endothermic reaction occurs, the heat required is absorbed from the thermal energy of the solution, which decreases its temperature (Figure below). The temperature change, along with the specific heat and mass of the solution, can then be used to calculate the amount of heat involved in either case.

In a calorimetric determination, either (a) an exothermic process occurs and heat, q, is negative, indicating that thermal energy is transferred from the system to its surroundings, or (b) an endothermic process occurs and heat, q, is positive, indicating that thermal energy is transferred from the surroundings to the system

Calorimetry measurements are important in understanding the heat transferred in reactions involving everything from microscopic proteins to massive machines .

Consider a simpler example that illustrates the core idea behind calorimetry. Suppose we initially have a high-temperature substance, such as a hot piece of metal (M), and a low-temperature substance, such as cool water (W). If we place the metal in the water, heat will flow from M to W. The temperature of M will decrease, and the temperature of W will increase, until the two substances have the same temperature—that is, when they reach thermal equilibrium (Figure below ). If this occurs in a calorimeter, ideally all of this heat transfer occurs between the two substances, with no heat gained  or lost by either its external environment

In a simple calorimetry process, (a) heat, q, is transferred from the hot metal, M, to the cool water, W, until (b) both are at the same temperature

Under these ideal circumstances, the net heat change is zero

This relationship can be rearranged to show that the heat gained by substance M is equal to the heat lost by substance W:

The magnitude of the heat (change) is therefore the same for both substances, and the negative sign merely shows that qsubstance M and qsubstance W are opposite in direction of heat flow (gain or loss) but does not indicate the arithmetic sign of either q value (that is determined by whether the matter in question gains or loses heat, per definition). In the specific situation described, qsubstance M is a negative value and qsubstance W is positive, since heat is transferred from M to W.

Read more

CHEMICAL KINETICS

From baking a cake to determining the useful lifespan of a bridge, rates of chemical reactions play important roles in our understanding of processes that involve chemical changes. Two questions are typically posed when planning to carry out a chemical reaction. The first is: “Will the reaction produce the desired products in useful quantities?” The second question is: “How rapidly will the reaction occur?” A third question is often asked when investigating reactions in greater detail: “What specific molecular-level processes take place as the reaction occurs?” Knowing the answer to this question is of practical importance when the yield or rate of a reaction needs to be controlled.

The study of chemical kinetics concerns the second and third questions—that is, the rate at which a reaction yields products and the molecular-scale means by which a reaction occurs

Chemical Reaction Rates

A rate is a measure of how some property varies with time. Speed is a familiar rate that expresses the distance traveled by an object in a given amount of time. Wage is a rate that represents the amount of money earned by a person working for a given amount of time. Likewise, the rate of a chemical reaction is a measure of how much reactant is consumed, or how much product is produced, by the reaction in a given amount of time.

The rate of reaction is the change in the amount of a reactant or product per unit time. Reaction rates are therefore determined by measuring the time dependence of some property that can be related to reactant or product amounts. Rates of reactions that consume or produce gaseous substances, for example, are conveniently determined by measuring changes in volume or pressure. For reactions involving one or more colored substances, rates may be monitored via measurements of light absorption. For reactions involving aqueous electrolytes, rates may be measured via changes in a solution’s conductivity

For reactants and products in solution, their relative amounts (concentrations) are conveniently used for purposes of expressing reaction rates. For example, the concentration of hydrogen peroxide, H2O2, in an aqueous solution changes slowly over time as it decomposes according to the equation

2H2O2 (aq) ———> 2H2O (l) + O2 (g)

The rate at which the hydrogen peroxide decomposes can be expressed in terms of the rate of change of its concentration, as shown here

This mathematical representation of the change in species concentration over time is the rate expression for the reaction. The brackets indicate molar concentrations, and the symbol delta (Δ) indicates “change in.” Thus [H2 O2]t1 represents the molar concentration of hydrogen peroxide at some time t1; likewise [H2 O2]t2  represents the molar concentration of hydrogen peroxide at some time t2 and Δ[H2O2] represents the change  in molar concentration of hydrogen peroxide during the time interval Δt (that is, t2t1). Since the reactant concentration decreases as the reaction proceeds, Δ[H2O2] is a negative quantity. Reaction rates are, by convention, positive quantities, and so this negative change in concentration is multiplied by −1. The table below provides an example of data collected during the decomposition of H2O2.

The rate of decomposition of H2O2 in an aqueous solution decreases as the concentration of H2O2  decreases.

To obtain the tabulated results for this decomposition, the concentration of hydrogen peroxide was measured every 6 hours over the course of a day at a constant temperature of 40 °C. Reaction rates were computed for each time interval by dividing the change in concentration by the corresponding time increment, as shown here for the first 6-hour period

This behavior indicates the reaction continually slows with time. Using the concentrations at the beginning and end of a time period over which the reaction rate is changing results in the calculation of an average rate for the reaction over this time interval. At any specific time, the rate at which a reaction is proceeding is known as its instantaneous rate. The instantaneous rate of a reaction at “time zero,” when the reaction commences, is its initial rate. Consider the analogy of a car slowing down as it approaches a stop sign. The vehicle’s initial rate—analogous to the beginning of a chemical reaction—would be the speedometer reading at the moment the driver begins pressing the brakes (t0). A few moments later, the instantaneous rate at a specific moment—call it t1—would be somewhat slower, as indicated by the speedometer reading at that point in time. As time passes, the instantaneous rate will continue to fall until it reaches zero, when the car (or reaction) stops. Unlike instantaneous speed, the car’s average speed is not indicated by the speedometer; but it can be calculated as the ratio of the distance traveled to the time required to bring the vehicle to a complete stop (Δt). Like the decelerating car, the average rate of a chemical reaction will fall somewhere between its initial and final rates.

The instantaneous rate of a reaction may be determined one of two ways. If experimental conditions permit the measurement of concentration changes over very short time intervals, then average rates computed as described earlier provide reasonably good approximations of instantaneous rates. Alternatively, a graphical procedure may be used that, in effect, yields the results that would be obtained if short time interval measurements were possible. In a plot of the concentration of hydrogen peroxide against time, the instantaneous rate of decomposition of H2O2 at any time t is given by the slope of a straight line that is tangent  to the curve at that time (Below Graph). These tangent line slopes may be evaluated using calculus

This graph shows a plot of concentration versus time for a 1.000 M solution of H2O2. The rate at any time is equal to the negative of the slope of a line tangent to the curve at that time. Tangents are shown at t = 0 h (“initial rate”) and at t = 12 h (“instantaneous rate” at 12 h)

Relative Rates of Reaction

The rate of a reaction may be expressed as the change in concentration of any reactant or product. For any given reaction, these rate expressions are all related simply to one another according to the reaction stoichiometry. The rate of the general reaction

aA —-à bB

can be expressed in terms of the decrease in the concentration of A or the increase in the concentration of B. These two rate expressions are related by the stoichiometry of the reaction:

Consider the reaction represented by the following equation:

2NH3 (g) —–à N2 (g) + 3 H2 (g)

The relation between the reaction rates expressed in terms of nitrogen production and ammonia consumption, for example, is:

Note that a negative sign has been included as a factor to account for the  opposite signs of the two amount changes (the reactant amount is decreasing while the product amount is increasing). For homogeneous reactions, both the reactants and products are present in the same solution and thus occupy the same volume, so the molar amounts may be replaced with molar concentrations

The below graph illustrates the change in concentrations over time for the decomposition of ammonia into nitrogen and hydrogen at 1100 °C. Slopes of the tangent lines at t = 500 s show that the instantaneous rates derived from all three species involved in the reaction are related by their stoichiometric factors. The rate of hydrogen production, for example, is observed to be three times greater than that for nitrogen production:

Changes in concentrations of the reactant and products for the reaction 2NH3 (g) —–à N2 (g) + 3 H2 (g)

 The  rates of change of the three concentrations are related by the reaction stoichiometry, as shown by the different slopes of the tangents at t = 500 s

Factors Affecting Reaction Rates

The rates at which reactants are consumed and products are formed during chemical reactions vary greatly. Five factors typically affecting the rates of chemical reactions will be explored in this section: the chemical nature of the reacting substances, the state of subdivision (one large lump versus many small particles) of the reactants, the temperature of the reactants, the concentration of the reactants, and the presence of a catalyst

The Chemical Nature of the Reacting Substances

The rate of a reaction depends on the nature of the participating substances. Reactions that appear similar may have different rates under the same conditions, depending on the identity of the reactants. For example, when small pieces of the metals iron and sodium are exposed to air, the sodium reacts completely with air overnight, whereas the iron is barely affected. The active metals calcium and sodium both react with water to form hydrogen gas and a base. Yet calcium reacts at a moderate rate, whereas sodium reacts so rapidly that the reaction is almost explosive

The Physical States of the Reactants

A chemical reaction between two or more substances requires intimate contact between the reactants. When reactants are in different physical states, or phases (solid, liquid, gaseous, dissolved), the reaction takes place only at the interface between the phases. Consider the heterogeneous reaction between a solid phase and either a liquid or gaseous phase. Compared with the reaction rate for large solid particles, the rate for smaller particles will be greater because the surface area in contact with the other reactant phase is greater. For example, large pieces of iron react more slowly with acids than they do with finely divided iron powder (Figure below ). Large pieces of wood smolder, smaller pieces burn rapidly, and saw dust burns explosively

(a) Iron powder reacts rapidly with dilute hydrochloric acid and produces bubbles of hydrogen gas: 2Fe(s) + 6HCl(aq) ———à 2FeCl3(aq) + 3H2(g). (b) An iron nail reacts more slowly because the surface area exposed to the acid is much less.

Temperature of the Reactants

Chemical reactions typically occur faster at higher temperatures. Food can spoil quickly when left on the kitchen counter. However, the lower temperature inside of a refrigerator slows that process so that the same food remains fresh for days. Gas burners, hot plates, and ovens are often used in the laboratory to increase the speed of reactions that proceed slowly at ordinary temperatures. For many chemical processes, reaction rates are approximately doubled when the temperature is raised by 10 °C.

Concentrations of the Reactants

The rates of many reactions depend on the concentrations of the reactants. Rates usually increase when the concentration of one or more of the reactants increases. For example, calcium carbonate (CaCO3) deteriorates as a result of its reaction with the pollutant sulfur dioxide. The rate of this reaction depends on the amount of sulfur dioxide in the air (Figure 17.7). An acidic oxide, sulfur dioxide combines with water vapor in the air to produce sulfurous acid in the following reaction :

In a polluted atmosphere where the concentration of sulfur dioxide is high, calcium carbonate deteriorates more rapidly than in less polluted air. Similarly, phosphorus burns much more rapidly in an atmosphere of pure oxygen than in air, which is only about 20% oxygen

Statues made from carbonate compounds such as limestone and marble typically weather slowly over time due to the actions of water, and thermal expansion and contraction. However, pollutants like sulfur dioxide can accelerate weathering. As the concentration of air pollutants increases, deterioration of limestone occurs more rapidly. (credit: James P Fisher III)

The Presence of a Catalyst

Relatively dilute aqueous solutions of hydrogen peroxide, H2O2, are commonly used as topical antiseptics. Hydrogen peroxide decomposes to yield water and oxygen gas according to the equation :

2 H2 O2 —-à 2 H2O (l) + O2 (g)

Under typical conditions, this decomposition occurs very slowly. When dilute H2O2(aq) is poured onto an open wound, however, the reaction occurs rapidly and the solution foams because of the vigorous production of oxygen gas. This dramatic difference is caused by the presence of substances within the wound’s exposed tissues that accelerate the decomposition process. Substances that function to increase the rate of a reaction are called catalysts .

Rate Laws

The rate of a reaction is often affected by the concentrations of reactants.  Rate laws (sometimes called differential rate laws) or rate equations are mathematical expressions that describe the relationship between the rate of a chemical reaction and the concentration of its reactants. As an example, consider the reaction described by the chemical equation

aA + bB ———- Products

where a and b are stoichiometric coefficients. The rate law for this reaction is written as:

 rate = k [ A ] m [ B ]

in which [A] and [B] represent the molar concentrations of reactants, and k is the rate constant, which is specific for a particular reaction at a particular temperature. The exponents m and n are the reaction orders and are typically positive integers, though they can be fractions, negative, or zero. The rate constant k and the reaction orders m and n must be determined experimentally by observing how the rate of a reaction changes as the concentrations of the reactants are changed. The rate constant k is independent of the reactant concentrations, but it does vary with temperature.

The reaction orders in a rate law describe the mathematical dependence of the rate on reactant concentrations. Referring to the generic rate law above, the reaction is m order with respect to A and n order with respect to B. For example, if m = 1 and n = 2, the reaction is first order in A and second order in B. The overall reaction order is simply the sum of orders for each reactant. For the example rate law here, the reaction is third order overall (1 + 2 = 3). A few specific examples are shown below to further illustrate this concept.

The rate law:

rate=k[H2O2]
describes a reaction that is first order in hydrogen peroxide and first order overall. The rate law:

rate=k[C4H6]2
describes a reaction that is second order in C4H6 and second order overall. The rate law:

rate=k[H+][OH]
describes a reaction that is first order in H+, first order in OH, and second order overall.

Reaction Order and Rate Constant Units

In some cases , the reaction orders in the rate law happen to be the same as the coefficients in the chemical equation for the reaction. This is merely a coincidence and very often not the case

Rate laws may exhibit fractional orders for some reactants, and negative reaction orders are sometimes observed when an increase in the concentration of one reactant causes a decrease in reaction rate. A few examples illustrating these points are provided

It is important to note that rate laws are determined by experiment only and are not reliably predicted by reaction stoichiometry

Note that the units in this table were derived using specific units for concentration (mol/L) and time (s), though any valid units for these two properties may be used.

Integrated Rate Laws

The rate laws discussed thus far relate the rate and the concentrations of reactants. We can also determine a second form of each rate law that relates the concentrations of reactants and time. These are called integrated rate laws. We can use an integrated rate law to determine the amount of reactant or product present after a period of time or to estimate the time required for a reaction to proceed to a certain extent. For example, an integrated rate law is used to determine the length of time a radioactive material must be stored for its radioactivity to decay to a safe level.

Using calculus, the differential rate law for a chemical reaction can be integrated with respect to time to give an equation that relates the amount of reactant or product present in a reaction mixture to the elapsed time of the reaction. This process can either be very straightforward or very complex, depending on the complexity of the differential rate law. For purposes of discussion, we will focus on the resulting integrated rate laws for first-

, second-, and zero-order reactions

First-Order Reactions

Integration of the rate law for a simple first-order reaction (rate = k[A]) results in an equation describing how the reactant concentration varies with time

[A]t = [A]0 e-kt

where [A]t is the concentration of A at any time t, [A]0 is the initial concentration of A, and k is the first-order  rate constant

For mathematical convenience, this equation may be rearranged to other formats, including direct and indirect proportionalities :

Second-Order Reactions

The equations that relate the concentrations of reactants and the rate constant of second-order reactions can be fairly complicated. To illustrate the point with minimal complexity, only the simplest second-order reactions will be described here, namely, those whose rates depend on the concentration of just one reactant. For these types of reactions, the differential rate law is written as:

rate = k [A]2

The integrated rate law for second-order reactions has the form of the equation of a straight line:

Zero-Order Reactions

For zero-order reactions, the differential rate law is:

                        rate = k

A zero-order reaction thus exhibits a constant reaction rate, regardless of the concentration of its reactant(s). This may seem counterintuitive, since the reaction rate certainly can’t be finite when the reactant concentration is zero. For purposes of this introductory text, it will suffice to note that zero-order kinetics are observed for some reactions only under certain specific conditions. These same reactions exhibit different kinetic behaviors when the specific conditions aren’t met, and for this reason the more prudent term pseudozero-order is sometimes used. The integrated rate law for a zero-order reaction is a linear function:

[A]t = -kt = [A]0

y = mx + b

A plot of [A] versus t for a zero-order reaction is a straight line with a slope of −k and a y-intercept of [A]0. Below graph  shows a plot of [NH3] versus t for the thermal decomposition of ammonia at the surface of two different heated solids. The decomposition reaction exhibits first-order behavior at a quartz (SiO2) surface, as suggested by the exponentially decaying plot of concentration versus time. On a tungsten surface, however, the plot is linear, indicating zero-order kinetics.

The decomposition of NH3 on a tungsten (W) surface is a zero-order reaction, whereas on a quartz (SiO2) surface, the reaction is first order

Read more

Basic Concepts of Chemistry – II

Atomic Structure and Symbolism

The development of modern atomic theory revealed much about the inner structure of atoms. It was learned that an atom contains a very small nucleus composed of positively charged protons and uncharged neutrons, surrounded by a much larger volume of space containing negatively charged electrons. The nucleus contains the majority of an atom’s mass because protons and neutrons are much heavier than electrons, whereas electrons occupy almost all of an atom’s volume. The diameter of an atom is on the order of 10−10 m, whereas the diameter of the nucleus is roughly 10−15 m—about 100,000 times smaller. For a perspective about their relative sizes, consider this: If the nucleus were the size of a blueberry, the atom would be about the size of a football stadium

Atoms—and the protons, neutrons, and electrons that compose them—are extremely small. For example, a carbon atom weighs less than 2 X 10-23 g ,  and an electron has a charge of less than 2 X 10-19 C (coulomb). When describing the properties of tiny objects such as atoms, we use appropriately small units of measure, such as the atomic mass unit (amu) and the fundamental unit of charge (e). The amu was originally defined based on hydrogen, the lightest element, then later in terms of oxygen. Since 1961, it has been defined with regard to the most abundant isotope of carbon, atoms of which are assigned masses of exactly 12 amu. (This isotope is known as “carbon-12” as will be discussed later in this module.) Thus, one amu is exactly of the mass of one carbon-12 atom ie 1 amu = 1.6605 X 10-24 g . (The Dalton (Da) and the unified atomic mass unit (u) are alternative units that are equivalent to the amu.) The fundamental unit of charge (also called the elementary charge) equals the magnitude of the charge of an electron (e) with e = 1.602 X 10-19 Coulomb

A proton has a mass of 1.0073 amu and a charge of 1+. A neutron is a slightly heavier particle with a mass 1.0087 amu and a charge of zero; as its name suggests, it is neutral. The electron has a charge of 1− and is a much lighter particle with a mass of about 0.00055 amu (it would take about 1800 electrons to equal the mass of one proton). The properties of these fundamental particles are summarized in the below table . ( You might notice that the sum of an atom’s subatomic particles does not equal the atom’s actual mass: The total mass of six protons, six neutrons, and six electrons is 12.0993 amu, slightly larger than 12.00 amu. This “missing” mass is known as the mass defect, and you will learn about it in  nuclear chemistry.)

Properties of Subatomic Particles

Name Location Charge (C) Unit Charge Mass (amu) Mass (g)
electron outside nucleus −1.602 10−19 1− 0.00055 0.00091 10−24
proton nucleus 1.602 10−19 1+ 1.00727 1.67262 10−24
neutron nucleus 0 0 1.00866 1.67493 10−24

The number of protons in the nucleus of an atom is its atomic number (Z). This is the defining trait of an element: Its value determines the identity of the atom. For example, any atom that contains six protons is the element carbon and has the atomic number 6, regardless of how many neutrons or electrons it may have. A neutral atom must contain the same number of positive and negative charges, so the number of protons equals the number of electrons. Therefore, the atomic number also indicates the number of electrons in an atom. The total number of protons and neutrons in an atom is called its mass number (A). The number of neutrons is therefore the difference between the mass number and the atomic number:

Mass Number (A) = Number of Protons + Number of Neutrons

A – Z = Number of Neutrons

Atoms are electrically neutral if they contain the same number of positively charged protons and negatively charged electrons. When the numbers of these subatomic particles are not equal, the atom is electrically charged and is called an ion. The charge of an atom is defined as follows:

Atomic charge = number of protons − number of electrons

Atoms (and molecules) typically acquire charge by gaining or losing electrons. An atom that gains one or more electrons will exhibit a negative charge and is called an anion. Positively charged atoms called cations are formed when an atom loses one or more electrons. For example, a neutral sodium atom (Z = 11) has 11 electrons. If this atom loses one electron, it will become a cation with a 1+ charge (11 − 10 = 1+). A neutral oxygen atom (Z = 8) has eight electrons, and if it gains two electrons it will become an anion with a 2− charge (8 − 10 = 2−).

Chemical Symbols

A chemical symbol is an abbreviation that we use to indicate an element or an atom of an element. For example, the symbol for mercury is Hg (Figure 2.13). We use the same symbol to indicate one atom of mercury (microscopic domain) or to label a container of many atoms of the element mercury (macroscopic domain).

The symbol Hg represents the element mercury regardless of the amount; it could represent one  atom of mercury or a large amount of mercury

The symbols for several common elements and their atoms are listed in the below table . Some symbols are derived from the common name of the element; others are abbreviations of the name in another language. Most symbols have one or two letters, but three-letter symbols have been used to describe some elements that have atomic numbers greater than 112. To avoid confusion with other notations, only the first letter of a symbol is capitalized. For example, Co is the symbol for the element cobalt, but CO is the notation for the compound carbon monoxide, which contains atoms of the elements carbon (C) and oxygen (O). All known elements and their symbols are in the periodic table .

Some Common Elements and Their Symbols

Element  Symbol Element Symbol
aluminum Al iron Fe (from ferrum)
bromine Br lead Pb (from plumbum)
calcium Ca magnesium Mg
carbon C mercury Hg (from hydrargyrum)
chlorine Cl nitrogen N
chromium Cr oxygen O
cobalt Co potassium K (from kalium)
copper Cu (from cuprum) silicon Si
fluorine F silver Ag (from argentum)
gold Au (from aurum) sodium Na (from natrium)
helium He sulfur S
hydrogen H tin Sn (from stannum)
iodine I zinc Zn

Traditionally, the discoverer (or discoverers) of a new element names the element. However, until the name is recognized by the International Union of Pure and Applied Chemistry (IUPAC), the recommended name of the new element is based on the Latin word(s) for its atomic number. For example, element 106 was called unnilhexium (Unh), element 107 was called unnilseptium (Uns), and element 108 was called unniloctium (Uno) for several years. These elements are now named after scientists (or occasionally locations); for example, element 106 is now known as seaborgium (Sg) in honor of Glenn Seaborg, a Nobel Prize winner who was active in the discovery of several heavy elements. Element 109 was named in honor of Lise Meitner, who discovered nuclear fission, a phenomenon that would have world-changing impacts; Meitner also contributed to the discovery of some major isotopes, discussed immediately below

Isotopes

The symbol for a specific isotope of any element is written by placing the mass number as a superscript (above the letter) to the left of the element symbol . The atomic number is sometimes written as a subscript (below the letter)  preceding the symbol, but since this number defines the element’s identity, as does its symbol, it is often omitted. For example, magnesium exists as a mixture of three isotopes, each with an atomic number of 12 and with mass numbers of 24, 25, and 26, respectively. These isotopes can be identified as 24Mg, 25Mg, and 26Mg. These isotope symbols are read as “element, mass number” and can be symbolized consistent with this reading. For instance, 24Mg is read as “magnesium 24,” and can be written as “magnesium-24” or “Mg-24.” 25Mg is read as “magnesium 25,” and can be written as “magnesium-25” or “Mg-25.” All magnesium atoms have 12 protons in their nucleus. They differ only because a 24Mg atom has 12 neutrons in its nucleus, a 25Mg atom has 13 neutrons, and a 26Mg has 14 neutrons

The symbol for an atom indicates the element via its usual two-letter symbol, the mass number as a left superscript, the atomic number as a left subscript (sometimes omitted), and the charge as a right superscript

Information about the naturally occurring isotopes of elements with atomic numbers 1 through 10 is given in the below table . Note that in addition to standard names and symbols, the isotopes of hydrogen are often referred to using common names and accompanying symbols. Hydrogen-2, symbolized 2H, is also called deuterium and sometimes symbolized D. Hydrogen-3, symbolized 3H, is also called tritium and sometimes symbolized T

Nuclear Compositions of Atoms of the Very Light Elements

Element Symbol  Atomic Number Number of Protons Number of Neutrons Mass (amu) % Natural Abundance
        hydrogen   (protium)   1   1   0   1.0078   99.989
  (deuterium)   1   1   1   2.0141   0.0115
  (tritium)   1   1   2   3.01605   — (trace)
  helium   2 2 1 3.01603 0.00013
  2 2 2 4.0026 100
  lithium   3 3 3 6.0151 7.59
  3 3 4 7.0160 92.41
beryllium   4 4 5 9.0122 100
  boron   5 5 5 10.0129 19.9
  5 5 6 11.0093 80.1
    carbon   6 6 6 12.0000 98.89
  6 6 7 13.0034 1.11
  6 6 8 14.0032 — (trace)
  nitrogen   7 7 7 14.0031 99.63
  7 7 8 15.0001 0.37
    oxygen   8 8 8 15.9949 99.757
  8 8 9 16.9991 0.038
  8 8 10 17.9992 0.205
fluorine   9 9 10 18.9984 100
neon   10 10 10 19.9924 90.48

Atomic Mass

Because each proton and each neutron contribute approximately one amu to the mass of an atom, and each electron contributes far less, the atomic mass of a single atom is approximately equal to its mass number (a whole number). However, the average masses of atoms of most elements are not whole numbers because most elements exist naturally as mixtures of two or more isotopes.

The mass of an element shown in a periodic table or listed in a table of atomic masses is a weighted, average mass of all the isotopes present in a naturally occurring sample of that element. This is equal to the sum of each individual isotope’s mass multiplied by its fractional abundance

For example, the element boron is composed of two isotopes: About 19.9% of all boron atoms are 10B with a mass of 10.0129 amu, and the remaining 80.1% are 11B with a mass of 11.0093 amu. The average atomic mass for boron is calculated to be:

Boron Average Mass =  (0.199 X 10.0129 amu) + (0.801 X 11.0093 amu)
= 11.99 amu + 8.82 amu
= 10.81 amu

It is important to understand that no single boron atom weighs exactly 10.8 amu; 10.8 amu is the average mass of all boron atoms, and individual boron atoms weigh either approximately 10 amu or 11 amu.

Chemical Formulas

A molecular formula is a representation of a molecule that uses chemical symbols to indicate the types of atoms followed by subscripts to show the number of atoms of each type in the molecule. (A subscript is used only when more than one atom of a given type is present.) Molecular formulas are also used as abbreviations for the names of compounds

The structural formula for a compound gives the same information as its molecular formula (the types and numbers of atoms in the molecule) but also shows how the atoms are connected in the molecule. The structural formula for methane contains symbols for one C atom and four H atoms, indicating the number of atoms in the molecule  (Below figure) . The lines represent bonds that hold the atoms together. (A chemical bond  is an attraction between atoms or ions that holds them together in a molecule or a crystal.  For now, simply know that the lines are an indication of how the atoms are connected in a molecule. A ball-and-stick model shows the geometric arrangement of the atoms with atomic sizes not to scale, and a space-filling model shows the relative sizes of the atoms

A methane molecule can be represented as (a) a molecular formula, (b) a structural formula, (c) a ball-and-stick model, and (d) a space-filling model. Carbon and hydrogen atoms are represented by black and white spheres, respectively

Although many elements consist of discrete, individual atoms, some exist as molecules made up of two or more atoms of the element chemically bonded together. For example, most samples of the elements hydrogen, oxygen, and nitrogen are composed of molecules that contain two atoms each (called diatomic molecules) and thus have the molecular formulas H2, O2, and N2, respectively. Other elements commonly found as diatomic molecules are fluorine (F2), chlorine (Cl2), bromine (Br2), and iodine (I2). The most common form of the element sulfur is composed of molecules that consist of eight atoms of sulfur; its molecular formula is S8

A molecule of sulfur is composed of eight sulfur atoms and is therefore written as S8. It can be represented as (a) a structural formula, (b) a ball-and-stick model, and (c) a space-filling model. Sulfur atoms are represented by yellow spheres

It is important to note that a subscript following a symbol and a number in front of a symbol do not represent the same thing; for example, H2 and 2H represent distinctly different species. H2 is a molecular formula; it represents a diatomic molecule of hydrogen, consisting of two atoms of the element that are chemically bonded together. The expression 2H, on the other hand, indicates two separate hydrogen atoms that are not combined as a unit. The expression 2H2 represents two molecules of diatomic hydrogen

The symbols H, 2H, H2, and 2H2 represent very different entities

Compounds are formed when two or more elements chemically combine, resulting in the formation of bonds. For example, hydrogen and oxygen can react to form water, and sodium and chlorine can react to form table salt. We sometimes describe the composition of these compounds with an empirical formula, which indicates the types of atoms present and the simplest whole-number ratio of the number of atoms (or ions) in the compound. For example, titanium dioxide (used as pigment in white paint and in the thick, white, blocking type of sunscreen) has an empirical formula of TiO2. This identifies the elements titanium (Ti) and oxygen (O) as the constituents of titanium dioxide, and indicates the presence of twice as many atoms of the element oxygen as atoms of the element titanium

The white compound titanium dioxide provides effective protection from the sun. (b) A crystal of titanium dioxide, TiO2, contains titanium and oxygen in a ratio of 1 to 2. The titanium atoms are gray and the oxygen atoms are red. (credit a: modification of work by “osseous”/Flickr)

As discussed previously, we can describe a compound with a molecular formula, in which the subscripts indicate the actual numbers of atoms of each element in a molecule of the compound. In many cases, the molecular formula of a substance is derived from experimental determination of both its empirical formula and its molecular mass (the sum of atomic masses for all atoms composing the molecule). For example, it can be determined experimentally that benzene contains two elements, carbon (C) and hydrogen (H), and that for every carbon atom in benzene, there is one hydrogen atom. Thus, the empirical formula is CH. An experimental determination of the molecular mass reveals that a molecule of benzene contains six carbon atoms and six hydrogen atoms, so the molecular formula for benzene is C6H6

Benzene, C6H6, is produced during oil refining and has many industrial uses. A benzene molecule can be represented as (a) a structural formula, (b) a ball-and-stick model, and (c) a space-filling model. (d) Benzene is a clear liquid. (credit d: modification of work by Sahar Atwa)

For example, the molecular formula for acetic acid, the component that gives vinegar its sharp taste, is C2H4O2. This formula indicates that a molecule of acetic acid  (Below Figure) contains two carbon atoms, four hydrogen atoms, and two oxygen atoms. The ratio of atoms is 2:4:2. Dividing by the lowest common denominator (2) gives the simplest, whole-number ratio of atoms, 1:2:1, so the empirical formula is CH2O. Note that a molecular formula is always a whole- number multiple of an empirical formula.

  • Vinegar contains acetic acid, C2H4O2, which has an empirical formula of CH2O. It can be represented as (b) a structural formula and (c) as a ball-and-stick model. (credit a: modification of work by “HomeSpot HQ”/Flickr)

It is important to be aware that it may be possible for the same atoms to be arranged in different ways: Compounds with the same molecular formula may have different atom-to-atom bonding and therefore different structures. For example, could there be another compound with the same formula as acetic acid, C2H4O2? And if so, what would be the structure of its molecules?

Two C atoms, four H atoms, and two O atoms can also be arranged to form a methyl formate, which is used in manufacturing, as an insecticide, and for quick-drying finishes. Methyl formate molecules have one of the oxygen atoms between the two carbon atoms, differing from the arrangement in acetic acid molecules. Acetic acid and methyl formate are examples of isomers—compounds with the same chemical formula but different molecular structures (Below Figure) . Note that this small difference in the arrangement of the atoms has a major effect on their respective chemical properties. You would certainly not want to use a solution of methyl formate as a substitute for a solution of acetic acid (vinegar) when you make salad dressing.

Molecules of (a) acetic acid and methyl formate (b) are structural isomers; they have the same formula (C2H4O2) but different structures (and therefore different chemical properties).

Formula Mass and the Mole Concept

Many argue that modern chemical science began when scientists started exploring the quantitative as well as the qualitative aspects of chemistry. For example, Dalton’s atomic theory was an attempt to explain the results of measurements that allowed him to calculate the relative masses of elements combined in various compounds. Understanding the relationship between the masses of atoms and the chemical formulas of compounds allows us to quantitatively describe the composition of substances.

Formula Mass

We have seen the development of the atomic mass unit, the concept of average atomic masses, and the use of chemical formulas to represent the elemental makeup of substances. These ideas can be extended to calculate the formula mass of a substance by summing the average atomic masses of  all the atoms represented in the substance’s formula.

Formula Mass for Covalent Substances

For covalent substances, the formula represents the numbers and types of atoms composing a single molecule of the substance; therefore, the formula mass may be correctly referred to as a molecular mass. Consider chloroform (CHCl3), a covalent compound once used as a surgical anesthetic and now primarily used in the production of tetrafluoroethylene, the building block for the “anti-stick” polymer, Teflon. The molecular formula of chloroform indicates that a single molecule contains one carbon atom, one hydrogen atom, and three chlorine atoms. The average molecular mass of a chloroform molecule is therefore equal to the sum of the average atomic masses of these atoms. Below figure outlines the calculations used to derive the molecular mass of chloroform, which is 119.37 amu

The average mass of a chloroform molecule, CHCl3, is 119.37 amu, which is the sum of the average atomic masses of each of its constituent atoms. The model shows the molecular structure of chloroform

Likewise, the molecular mass of an aspirin molecule, C9H8O4, is the sum of the atomic masses of nine carbon atoms, eight hydrogen atoms, and four oxygen atoms, which amounts to 180.15 amu

The average mass of an aspirin molecule is 180.15 amu. The model shows the molecular structure of aspirin, C9H8O4

Formula Mass for Ionic Compounds

Ionic compounds are composed of discrete cations and anions combined in ratios to yield electrically neutral bulk matter. The formula mass for an ionic compound is calculated in the same way as the formula mass for covalent compounds: by summing the average atomic masses of all the atoms in the compound’s formula. Keep in mind, however, that the formula for an ionic compound does not represent the composition of a  discrete molecule, so it may not correctly be referred to as the “molecular mass.”

As an example, consider sodium chloride, NaCl, the chemical name for common table salt. Sodium chloride is an ionic compound composed of sodium cations, Na+, and chloride anions, Cl−, combined in a 1:1 ratio. The formula mass for this compound is computed as 58.44 amu

Table salt, NaCl, contains an array of sodium and chloride ions combined in a 1:1 ratio. Its formula mass is 58.44 amu

Note that the average masses of neutral sodium and chlorine atoms were used in this computation, rather than   the masses for sodium cations and chlorine anions. This approach is perfectly acceptable when computing the formula mass of an ionic compound. Even though a sodium cation has a slightly smaller mass than a sodium atom (since it is missing an electron), this difference will be offset by the fact that a chloride anion is slightly more massive than a chloride atom (due to the extra electron). Moreover, the mass of an electron is negligibly small with respect to the mass of a typical atom. Even when calculating the mass of an isolated ion, the missing or additional electrons can generally be ignored, since their contribution to the overall mass is negligible, reflected only in the nonsignificant digits that will be lost when the computed mass is properly rounded. The few exceptions to this guideline are very light ions derived from elements with precisely known atomic masses.

The Mole Concept and Molar Mass

The identity of a substance is defined not only by the types of atoms or ions it contains, but by the quantity of each type of atom or ion. For example, water, H2O, and hydrogen peroxide, H2O2, are alike in that their respective molecules are composed of hydrogen and oxygen atoms. However, because a hydrogen peroxide molecule contains two oxygen atoms, as opposed to the water molecule, which has only one, the two substances exhibit very different properties. Today, sophisticated instruments allow the direct measurement of these defining microscopic traits; however, the same traits were originally derived from the measurement of macroscopic properties (the masses and volumes of bulk quantities of matter) using relatively simple tools (balances and volumetric glassware). This experimental approach required the introduction of a new unit for amount of substances, the mole, which remains indispensable in modern chemical science.

The mole is an amount unit similar to familiar units like pair, dozen, gross, etc. It provides a specific measure of the number of atoms or molecules in a sample of matter. One Latin connotation for the word “mole” is “large mass” or “bulk,” which is consistent with its use as the name for this unit. The mole provides a link between an easily measured macroscopic property, bulk mass, and an  extremely  important  fundamental  property, number of atoms, molecules, and so forth. A mole of substance is that amount in which there are 6.02214076 X 1023 discrete entities (atoms or molecules). This large number is a fundamental constant known as Avogadro’s number (NA) or the Avogadro constant in honor of Italian scientist Amedeo Avogadro. This constant is properly reported with an explicit unit of “per mole,” a conveniently rounded version being 6.022 X 1023 / mol

Consistent with its definition as an amount unit, 1 mole of any element contains the same number of atoms as 1 mole of any other element. The masses of 1 mole of different elements, however, are different, since the masses of the individual atoms are drastically different. The molar mass of an element (or compound) is the mass in grams of 1 mole of that substance, a property expressed in units of grams per mole (g/mol)

Each sample contains 6.022 X 10 23  atoms —1.00 mol of atoms. From left to right (top row): 65.4 g zinc, 12.0 g carbon, 24.3 g magnesium, and 63.5 g copper. From left to right (bottom row): 32.1 g sulfur, 28.1 g silicon, 207 g lead, and 118.7 g tin. (credit: modification of work by Mark Ott)

The molar mass of any substance is numerically equivalent to its atomic or formula weight in amu. Per the amu definition, a single 12C atom weighs 12 amu (its atomic mass is 12 amu). A mole of 12C weighs 12 g (its molar mass is 12 g/mol). This relationship holds for all elements, since their atomic masses are measured relative to that of the amu-reference substance, 12C. Extending this principle, the molar mass of a compound in   grams is likewise numerically equivalent to its formula mass in amu

Each sample contains 6.02 X 1023  molecules or formula units—1.00 mol of the compound or element. Clock-wise from the upper left: 130.2 g of C8H17OH (1-octanol, formula mass 130.2 amu), 454.4 g of HgI2 (mercury(II) iodide, formula mass 454.4 amu), 32.0 g of CH3OH (methanol, formula mass 32.0 amu) and 256.5 g of S8 (sulfur, formula mass 256.5 amu). (credit: Sahar Atwa)

While atomic mass and molar mass are numerically equivalent, keep in mind that they are vastly different in terms of scale, as represented by the vast difference in the magnitudes of their respective units (amu versus g). To appreciate the enormity of the mole, consider a small drop of water weighing about 0.03 g . Although this represents just a tiny fraction of 1 mole of water (~18 g), it contains more water molecules than can be clearly imagined. If the molecules were distributed equally among the roughly seven billion people on earth, each person would receive more than 100 billion molecules.

Determining Empirical and Molecular Formulas

The previous section discussed the relationship between the bulk mass of a substance and the number of atoms or molecules it contains (moles). Given the chemical formula of the substance, one may determine the amount of the substance (moles) from its mass, and vice versa. But what if the chemical formula of a substance is unknown? In this section, these same principles will be applied to derive the chemical formulas of unknown substances from experimental mass measurements

Percent Composition

The elemental makeup of a compound defines its chemical identity, and chemical formulas are the most succinct way of representing this elemental makeup. When a compound’s formula is unknown, measuring the mass of each of its constituent elements is often the first step in the process of determining the formula experimentally. The results of these measurements permit the calculation of the compound’s percent composition, defined as the percentage by mass of each element in the compound.

For example, consider a gaseous compound composed solely of carbon and hydrogen. The percent composition of this compound could be represented as follows :

% H = ( Mass H / Mass Compound ) X 100%

% C = ( Mass C / Mass Compound ) X 100%

If analysis of a 10.0-g sample of this gas showed it to contain 2.5 g H and 7.5 g C, the percent composition would be calculated to be 25% H and 75% C:

% H = (0.25 g / 10 g ) X 100% = 25%

% C = (0.75 g / 10 g ) X 100% = 75%

Determining Percent Composition from Molecular or Empirical Formulas

Percent composition is also useful for evaluating the relative abundance of a given element in different compounds of known formulas. As one example, consider the common nitrogen-containing fertilizers ammonia (NH3), ammonium nitrate (NH4NO3), and urea (CH4N2O). The element nitrogen is the active ingredient for agricultural purposes, so the mass percentage of nitrogen in the compound is a practical and economic concern for consumers choosing among these fertilizers. For these sorts of applications, the percent composition of a compound is easily derived from its formula mass and the atomic masses of its constituent elements. A molecule of NH3 contains one N atom weighing 14.01 amu and three H atoms weighing a total of ( 3 X 1.008 amu) = 3.024 amu . The formula mass of ammonia is therefore (14.01 amu + 3.024 amu) = 17.03 amu, and its percent composition is:

% N = (14.01  / 17.03) X 100 = 82.2%

% H = (3.024 / 17.03) X 100 = 18.8 %

This same approach may be taken considering a pair of molecules, a dozen molecules, or a mole of molecules, etc. The latter amount is most convenient and would simply involve the use of molar masses instead of atomic and formula masses. As long as the molecular or empirical formula of the compound in question is known, the percent composition may be derived from the atomic or molar masses of the compound’s elements

Determination of Empirical Formulas :

The most common approach to determining a compound’s chemical formula is to first measure the masses of its constituent elements. However, keep in mind that chemical formulas represent the relative numbers, not masses, of atoms in the substance. Therefore, any experimentally derived data involving mass must be used to derive the corresponding numbers of atoms in the compound. This is accomplished using molar masses to convert the mass of each element to a number of moles. These molar amounts are used to compute whole-number ratios that can be used to derive the empirical formula of the substance

In summary, empirical formulas are derived from experimentally measured element masses by:

  1. Deriving the number of moles of each element from its mass
  2. Dividing each element’s molar amount by the smallest molar amount to yield subscripts for a tentative empirical formula
  3. Multiplying all coefficients by an integer, if necessary, to ensure that the smallest whole-number ratio of subscripts is obtained

Below Figure outlines this procedure in flow chart fashion for a substance containing elements A and X.

Finally, with regard to deriving empirical formulas, consider instances in which a compound’s percent composition is available rather than the absolute masses of the compound’s constituent elements. In such cases, the percent composition can be used to calculate the masses of elements present in any convenient mass of compound; these masses can then be used to derive the empirical formula in the usual fashion.

Derivation of Molecular Formulas

Recall that empirical formulas are symbols representing the relative numbers of a compound’s elements. Determining the absolute numbers of atoms that compose a single molecule of a covalent compound requires knowledge of both its empirical formula and its molecular mass or molar mass. These quantities may be determined experimentally by various measurement techniques. Molecular mass, for example, is often derived from the mass spectrum of the compound  . Molar mass can be measured by a number of experimental methods

Molecular formulas are derived by comparing the compound’s molecular or molar mass to its empirical formula mass. As the name suggests, an empirical formula mass is the sum of the average atomic masses of all the atoms represented in an empirical formula. If the molecular (or molar) mass of the substance is known, it may be divided by the empirical formula mass to yield the number of empirical formula units per molecule (n)

Molarity

Mixtures—samples of matter containing two or more substances physically combined—are more commonly encountered in nature than are pure substances.

Similar to a pure substance, the relative composition of a mixture plays an important role in determining its properties. The relative amount of oxygen in a planet’s atmosphere determines its ability to sustain aerobic life. The relative amounts of iron, carbon, nickel, and other elements in steel (a mixture known as an “alloy”) determine its physical strength and resistance to corrosion. The relative amount of the active ingredient in a medicine determines its effectiveness in achieving the desired pharmacological effect. The relative amount of sugar in a beverage determines its sweetness . This section will describe one of the most common ways in which the relative compositions of mixtures may be quantified.

Sugar is one of many components in the complex mixture known as coffee. The amount of sugar in a given amount of coffee is an important determinant of the beverage’s sweetness. (credit: Jane Whitney)

Solutions

Solutions have previously been defined as homogeneous mixtures, meaning that the composition of the mixture (and therefore its properties) is uniform throughout its entire volume. Solutions occur frequently in nature and have also been implemented in many forms of manmade technology. A more thorough treatment of solution properties is provided in the chapter on solutions and colloids, but provided here is an introduction to some of the basic properties of solutions.

The relative amount of a given solution component is known as its concentration. Often, though not always, a solution contains one component with a concentration that is significantly greater than that of all other components. This component is called the solvent and may be viewed as the medium in which the other components are dispersed, or dissolved. Solutions in which water is the solvent are, of course, very common on our planet. A solution in which water is the solvent is called an aqueous solution.

A solute is a component of a solution that is typically present at a much lower concentration than the solvent. Solute concentrations are often described with qualitative terms such as dilute (of relatively low concentration) and concentrated (of relatively high concentration).

Concentrations may be quantitatively assessed using a wide variety of measurement units, each convenient for particular applications. Molarity (M) is a useful concentration unit for many applications in chemistry.

Molarity is defined as the number of moles of solute in exactly 1 liter (1 L) of the solution:

ie  [ M = ( mol Solute / L solution) ]

Mass Percentage

Percentages are also commonly used to express the composition of mixtures, including solutions. The mass percentage of a solution component is defined as the ratio of the component’s mass to the solution’s mass, expressed as a percentage

Mass Percentage = (Mass of Component / Mass of Solution) X 100 %

Mass percentage is also referred to by similar names such as percent mass, percent weight, weight/weight percent, and other variations on this theme. The most common symbol for mass percentage is simply the percent sign, %, although more detailed symbols are often used including %mass, %weight, and (w/w)%. Use of these more detailed symbols can prevent confusion of mass percentages with other types of percentages, such as volume percentages .

Mass percentages are popular concentration units for consumer products. The label of a typical liquid bleach bottle cites the concentration of its active ingredient, sodium hypochlorite (NaOCl), as being 7.4%. A 100.0-g sample of bleach would therefore contain 7.4 g of NaOCl

Liquid bleach is an aqueous solution of sodium hypochlorite (NaOCl). This brand has a concentration of 7.4% NaOCl by mass

Volume Percentage

Liquid volumes over a wide range of magnitudes are conveniently measured using common and relatively inexpensive laboratory equipment. The concentration of a solution formed by dissolving a liquid solute in a liquid solvent is therefore often expressed as a volume percentage, %vol or (v/v)%

Volume Percentage = (Volume Solute / Volume Solution) X 100 %

Mass-Volume Percentage

“Mixed” percentage units, derived from the mass of solute and the volume of solution, are popular for certain biochemical and medical applications. A mass-volume percent is a ratio of a solute’s mass to the solution’s volume expressed as a percentage. The specific units used for solute mass and solution volume may vary, depending on the solution. For example, physiological saline solution, used to prepare intravenous fluids, has a concentration of 0.9% mass/volume (m/v), indicating that the composition is 0.9 g of solute per 100 mL of solution. The concentration of glucose in blood (commonly referred to as “blood sugar”) is also typically expressed in terms of a mass-volume ratio. Though not expressed explicitly as a percentage, its concentration is usually given in milligrams of glucose per deciliter (100 mL) of blood

“Mixed” mass-volume units are commonly encountered in medical settings. (a) The NaCl concentration of physiological saline is 0.9% (m/v). (b) This device measures glucose levels in a sample of blood. The normal range for glucose concentration in blood (fasting) is around 70–100 mg/dL. (credit a: modification of work by “The National Guard”/Flickr; credit b: modification of work by Biswarup Ganguly)

Parts per Million and Parts per Billion

Very low solute concentrations are often expressed using appropriately small units such as parts per million (ppm) or parts per billion (ppb). Like percentage (“part per hundred”) units, ppm and ppb may be defined in terms of masses, volumes, or mixed mass-volume units. There are also ppm and ppb units defined with respect to numbers of atoms and molecules

Both ppm and ppb are convenient units for reporting the concentrations of pollutants and other trace contaminants in water. Concentrations of these contaminants are typically very low in treated and natural waters, and their levels cannot exceed relatively low concentration thresholds without causing adverse effects on health and wildlife. For example, the EPA has identified the maximum safe level of fluoride ion in tap water to be 4 ppm. Inline water filters are designed to reduce the concentration of fluoride and several other trace- level contaminants in tap water

  • In some areas, trace-level concentrations of contaminants can render unfiltered tap water unsafe for drinking and cooking. (b) Inline water filters reduce the concentration of solutes in tap water. (credit a: modification of work by Jenn Durfey; credit b: modification of work by “vastateparkstaff”/Wikimedia commons)

Stoichiometry of Chemical Reactions

The description of  how to symbolize chemical reactions using chemical equations, how   to classify some common chemical reactions by identifying patterns of reactivity, and how to determine the quantitative relations between the amounts of substances involved in chemical reaction is called  the reaction stoichiometry.

Writing and Balancing Chemical Equations

When atoms gain or lose electrons to yield ions, or combine with other atoms to form molecules, their symbols are modified or combined to generate chemical formulas that appropriately represent these species. Extending this symbolism to represent both the identities and the relative quantities of substances undergoing a chemical (or physical) change involves writing and balancing a chemical equation.

Consider as an example the reaction between one methane molecule (CH4) and two diatomic oxygen molecules (O2) to produce one carbon dioxide molecule (CO2) and two water molecules (H2O). The chemical equation representing this process is provided in the upper half of the below figure, with space-filling molecular models shown in the lower half of the figure.

The reaction between methane and oxygen to yield carbon dioxide and water (shown at bottom) may be represented by a chemical equation using formulas (top).

This example illustrates the fundamental aspects of any chemical equation:

  1. The substances undergoing reaction are called reactants, and their formulas are placed on the left side of the equation.
  2. The substances generated by the reaction are called products, and their formulas are placed on the right side of the equation.
  3. Plus signs (+) separate individual reactant and product formulas, and an arrow -> separates the reactant and product (left and right) sides of the equation.
  4. The relative numbers of reactant and product species are represented by coefficients (numbers placed immediately to the left of each formula). A coefficient of 1 is typically omitted.

It is common practice to use the smallest possible whole-number coefficients in a chemical equation, as is done in this example. Realize, however, that these coefficients represent the relative numbers of reactants and products, and, therefore, they may be correctly interpreted as ratios. Methane and oxygen react to yield carbon dioxide and water in a 1:2:1:2 ratio. This ratio is satisfied if the numbers of these molecules are, respectively, 1-2-1-2 , or 2-4-2-4, or 3-6-3-6, and so on . Likewise, these coefficients may be interpreted with regard to any amount (number) unit, and so this equation may be correctly read in many ways, including:

  • One methane molecule and two oxygen molecules react to yield one carbon dioxide molecule and two water molecules.
  • One dozen methane molecules and two dozen oxygen molecules react to yield one dozen carbon dioxide molecules and two dozen water molecules.
  • One mole of methane molecules and 2 moles of oxygen molecules react to yield 1 mole of carbon dioxide molecules and 2 moles of water molecules

Regardless of the absolute numbers of molecules involved, the ratios between numbers of molecules of each species that react (the reactants) and molecules of each species that form (the products) are the same and are given by the chemical reaction equation

Balancing Equations

A chemical equation is balanced, if  equal numbers of atoms for each element involved in the reaction are represented on the reactant and product sides. This is a requirement the equation must satisfy to be consistent with the law of conservation of matter. It may be confirmed by simply summing the numbers of atoms on either side of the arrow and comparing these sums to ensure they are equal. Note that the number of atoms for a given element is calculated by multiplying the coefficient of any formula containing that element by the element’s subscript in the formula. If an element appears in more than one formula on a given side of the equation, the number of atoms represented in each must be computed and then added together

A balanced chemical equation provides a great deal of information in a very succinct format. Chemical formulas provide the identities of the reactants and products involved in the chemical change, allowing classification of the reaction. Coefficients provide the relative numbers of these chemical species, allowing a quantitative assessment of the relationships between the amounts of substances consumed and produced by the reaction. These quantitative relationships are known as the reaction’s stoichiometry, a term derived from the Greek words stoicheion (meaning “element”) and metron (meaning “measure”). In this module, the use of balanced chemical equations for various stoichiometric applications is explored

The general approach to using stoichiometric relationships is similar in concept to the way people go about many common activities. Food preparation, for example, offers an appropriate comparison. A recipe for making eight pancakes calls for 1 cup pancake mix, ¾ cup milk , and one egg. The “equation” representing the  preparation of pancakes per this recipe is

1 cup pancake mix + ¾ cup milk + 1 egg ——> 8 pancakes

If two dozen pancakes are needed for a big family breakfast, the ingredient amounts must be increased proportionally according to the amounts given in the recipe. For example, the number of eggs required to make 24 pancakes is

24 pancakes X ( 1 egg / 8 pancakes) = 3 eggs

Balanced chemical equations are used in much the same fashion to determine the amount of one reactant required to react with a given amount of another reactant, or to yield a given amount of product, and so forth. The coefficients in the balanced equation are used to derive stoichiometric factors that permit computation of the desired quantity

Numerous variations on the beginning and ending computational steps are possible depending upon what particular quantities are provided and sought (volumes, solution concentrations, and so forth). Regardless of the details, all these calculations share a common essential component: the use of stoichiometric factors derived from balanced chemical equations. Below Figure  provides a general outline of the various computational steps associated with many reaction stoichiometry calculations.

Reaction Yields

The relative amounts of reactants and products represented in a balanced chemical equation are often referred to as stoichiometric amounts. All the exercises of the preceding module involved stoichiometric amounts of reactants. For example, when calculating the amount of product generated from a given amount of reactant, it was assumed that any other reactants required were available in stoichiometric amounts (or greater).

Limiting Reactant

Consider another food analogy, making grilled cheese sandwiches

1 slice of cheese + 2 slices of Bread ———->1 sandwich

Stoichiometric amounts of sandwich ingredients for this recipe are bread and cheese slices in a 2:1 ratio

If you are Provided with 28 slices of bread and 11 slices of cheese, you may prepare 11 sandwiches per the provided recipe, using all the provided cheese and having six slices of bread left over. In this scenario, the number of sandwiches prepared has been limited by the number of cheese slices, and the bread slices have been provided in excess.

Consider this concept now with regard to a chemical process, the reaction of hydrogen with chlorine to yield  hydrogen chloride

H2​+Cl2 →2HCl

The balanced equation shows the hydrogen and chlorine react in a 1:1 stoichiometric ratio. If these reactants are provided in any other amounts, one of the reactants will nearly always be entirely consumed, thus limiting the amount of product that may be generated. This substance is the limiting reactant, and the other substance is the excess reactant. Identifying the limiting and excess reactants for a given situation requires computing the molar amounts of each reactant provided and comparing them to the stoichiometric amounts represented in the balanced chemical equation. For example, imagine combining 3 moles of H2 and 2 moles of Cl2. This represents a 3:2 (or 1.5:1) ratio of hydrogen to chlorine present for reaction, which is greater than the stoichiometric ratio of 1:1. Hydrogen, therefore, is present in excess, and chlorine is the limiting reactant.

Reaction of all the provided chlorine (2 mol) will consume 2 mol of the 3 mol of hydrogen provided, leaving 1 mol of hydrogen unreacted.

An alternative approach to identifying the limiting reactant involves comparing the amount of product expected for the complete reaction of each reactant. Each reactant amount is used to separately calculate the amount of product that would be formed per the reaction’s stoichiometry. The reactant yielding the lesser amount of product is the limiting reactant

Mole Fraction and Molality

Several units commonly used to express the concentrations of solution components have been seen ,  each providing certain benefits for use in different applications. For example, molarity (M) is a convenient unit for use in stoichiometric calculations, since it is defined in terms of the molar amounts of solute species

 [ M = ( mol Solute / L solution) ]

Because solution volumes vary with temperature, molar concentrations will likewise vary. When expressed as molarity, the concentration of a solution with identical numbers of solute and solvent species will be different at different temperatures, due to the contraction/expansion of the solution. More appropriate for calculations involving many colligative properties are mole-based concentration units whose values are not dependent on temperature. Two such units are mole fraction  and molality.

The mole fraction, X, of a component is the ratio of its molar amount to the total number of moles of all solution components . By this definition, the sum of mole fractions for all solution components (the solvent and all solutes) is equal to one.

Molality is a concentration unit defined as the ratio of the numbers of moles of solute to the mass of the solvent in kilograms

ie [ m = mol Solute / Kg Solvent]

Since these units are computed using only masses and molar amounts, they do not vary with temperature and, thus, are better suited for applications requiring temperature-independent concentrations, including several colligative properties

Take Quiz

1. Write the symbol for each of the following ions:

a) the ion with a 1+ charge, atomic number 55, and mass number 133
b) the ion with 54 electrons, 53 protons, and 74 neutrons
c) the ion with atomic number 15, mass number 31, and a 3− charge
d) the ion with 24 electrons, 30 neutrons, and a 3+ charge

ANSWER

a) The ion with a 1+ charge, atomic number 55, and mass number 133 is the cesium ion.
b) The ion with 54 electrons, 53 protons, and 74 neutrons is the iodide ion.
c) The ion with atomic number 15, mass number 31, and a 3− charge is the phosphide ion.
d) The ion with 24 electrons, 30 neutrons, and a 3+ charge is the cobalt(III) ion.

2. Determine the number of protons, neutrons, and electrons in the following isotopes that are used in medical diagnoses:
a) atomic number 9, mass number 18, charge of 1−
b) atomic number 43, mass number 99, charge of 7+
c) atomic number 53, atomic mass number 131, charge of 1−
d) atomic number 81, atomic mass number 201, charge of 1+

ANSWER

Let’s determine the number of protons, neutrons, and electrons for each given isotope:
a) **Atomic number 9, mass number 18, charge of 1−**
– Protons: 9 (since the atomic number is 9)
– Neutrons: 18 – 9 = 9 (mass number – atomic number)
– Electrons: 9 + 1 = 10 (since the charge is 1−, there is one extra electron)
**Summary:**
– Protons: 9
– Neutrons: 9
– Electrons: 10

b) **Atomic number 43, mass number 99, charge of 7+**
– Protons: 43 (since the atomic number is 43)
– Neutrons: 99 – 43 = 56 (mass number – atomic number)
– Electrons: 43 – 7 = 36 (since the charge is 7+, there are seven fewer electrons) **Summary:**
– Protons: 43
– Neutrons: 56
– Electrons: 36

c) **Atomic number 53, mass number 131, charge of 1−**
– Protons: 53 (since the atomic number is 53)
– Neutrons: 131 – 53 = 78 (mass number – atomic number)

– Electrons: 53 + 1 = 54 (since the charge is 1−, there is one extra electron) **Summary:**
– Protons: 53
– Neutrons: 78
– Electrons: 54

d) **Atomic number 81, mass number 201, charge of 1+**
– Protons: 81 (since the atomic number is 81)
– Neutrons: 201 – 81 = 120 (mass number – atomic number)
– Electrons: 81 – 1 = 80 (since the charge is 1+, there is one fewer electron) **Summary:**
– Protons: 81
– Neutrons: 120
– Electrons: 80

3. An element has the following natural abundances and isotopic masses: 90.92% abundance with 19.99 amu, 0.26% abundance with 20.99 amu, and 8.82% abundance with 21.99 amu. Calculate the average atomic mass of this element

ANSWER

The formula for the average atomic mass is:
Average Atomic Mass = ∑ (Fractional Abundance X Isotopic Mass)
Given:
– Isotope 1: 90.92% abundance with 19.99 amu
– Isotope 2: 0.26% abundance with 20.99 amu
– Isotope 3: 8.82% abundance with 21.99 amu
First, convert the percentages to decimal form (fractional abundances):
– Isotope 1: 90.92% = 0.9092
– Isotope 2: 0.26% = 0.0026
– Isotope 3: 8.82% = 0.0882

Now, calculate the average atomic mass:
Average Atomic Mass = (0.9092 X 19.99) + (0.0026 X 20.99) + (0.0882 X 21.99)
= (18.172108 + 0.054574 + 1.940718)
= 20.1674 amu

4. Average atomic masses listed by IUPAC are based on a study of experimental results. Bromine has two isotopes, 79Br and 81Br, whose masses (78.9183 and 80.9163 amu, respectively) and abundances (50.69% and 49.31%, respectively) were determined in earlier experiments. Calculate the average atomic mass of bromine based on these experiments.

ANSWER

79Br – Mass = 78.9183 amu and Abundance = 50.69% , which is 0.5069 in fractional form
81Br – mass = 80.9163 amu and Abundance = 49.31% , which is 0.4931 in fractional form
Average atomic mass= (0.5069×78.9183)+(0.4931×80.9163) = 79.8909
Hence , the average atomic mass = 79.90 amu

5. Determine the empirical formulas for the following compounds:

(a) caffeine, C8H10N4O2
(b) sucrose, C12H22O11
(c) hydrogen peroxide, H2O2
(d) glucose, C6H12O6
(e) ascorbic acid (vitamin C), C6H8O6

ANSWER

(a) C4H5N2O;
(b) C12H22O11;
(c) HO;
(d) CH2O;
(e) C3H4O3

6. Name the following compounds:

a) CsCl
b) BaO
c) K2 S
d) BeCl2
e) Hbr
f) AlF3

ANSWER

(a) cesium chloride;
(b) barium oxide;
(c) potassium sulfide;
(d) beryllium chloride;
(e) hydrogen bromide;
(f ) aluminum fluoride

7. Write the formulas of the following compounds:
(a) chlorine dioxide
(b) dinitrogen tetraoxide
(c) potassium phosphide
(d) silver(I) sulfide
(e) Aluminium fluoride trihydrate
(f) Silicon dioxide

ANSWER

(a) ClO2;
(b) N2O4;
(c) K3P;
(d) Ag2S;
(e) AIF3•3H2O;
(f ) SiO2

8. Calculate the molecular or formula mass of each of the following:

(a) P4
(b) H2O
(c) Ca(NO3)2
(d) CH3CO2H (acetic acid)
(e) C12H22O11 (sucrose, cane sugar)

ANSWER

(a) 123.896 amu;
(b) 18.015 amu;
(c) 164.086 amu;
(d) 60.052 amu;
(e) 342.297 amu

9. Calculate the molar mass of each of the following:
a) S8
b) C5H12
c) Sc2(SO4)3
d) CH3COCH3 (acetone)
e) C6H12O6 (glucose)

ANSWER

(a) 256.48 g/mol;
(b) 72.150 g mol−1;
(c) 378.103 g mol−1;
(d) 58.080 g mol−1;
(e) 180.158 g mol−1

10. Determine the mass of each of the following:
(a) 2.345 mol LiCl
(b) 0.0872 mol acetylene, C2H2
(c) 3.3 X 10-2 mol Na2CO3
(d)1.23 X 103 mol Fructose, C6H12O6
(e) 0.5758 mol FeSO4(H2O)7

ANSWER

(a) 99.41 g;
(b) 2.27 g;
(c) 3.5 g;
(d) 222 kg;
(e) 160.1 g

11. Determine which of the following contains the greatest mass of aluminum: 122 g of AlPO4, 266 g of Al2Cl6, or 225 g of Al2S3

ANSWER

AlPO4: 1.000 mol, or 26.98 g Al;
Al2Cl6: 1.994 mol, or 53.74 g Al;
Al2S3: 3.00 mol, or 80.94 g Al;
Hence , The Al2S3 sample thus contains the greatest mass of Al.

12. A compound of carbon and hydrogen contains 92.3% C and has a molar mass of 78.1 g/mol. What is its molecular formula ?

ANSWER

C6H6

13. Calculate the number of moles and the mass of the solute in each of the following solutions:
a) 2.00 L of 18.5 M H2SO4, concentrated sulfuric acid
b) 100.0 mL of 3.8 X 10-6 M NaCN , the minimum lethal concentration of sodium cyanide in blood serum
c) 5.50 L of 13.3 M H2CO, the formaldehyde used to “fix” tissue samples
d) 325 mL of 1.8 X 10-6 M FeSO4, the minimum concentration of iron sulfate detectable by taste in drinking water

ANSWER

a) 37.0 mol H2SO4, , 3.63 X 103 g H2SO4
b) 3.8 X 10-7 mol NaCN , 1.9 X 10-5 g NaCN
c) 73.2 mol H2CO,2.20 kg H2CO
d) 5.9 X 10-7 mol FeSO4, 8.9 X 10-5 g FeSO4

14. Calculate the molarity of each of the following solutions:
a) 0.195 g of cholesterol, C27H46O, in 0.100 L of serum, the average concentration of cholesterol in human serum
b) 4.25 g of NH3 in 0.500 L of solution, the concentration of NH3 in household ammonia
c) 1.49 kg of isopropyl alcohol, C3H7OH, in 2.50 L of solution, the concentration of isopropyl alcohol in rubbing alcohol
d) 0.029 g of I2 in 0.100 L of solution, the solubility of I2 in water at 20 °C

ANSWER

(a) 5.04 X 10-2 M;
(b) 0.499 M;
(c) 9.92 M;
(d) 1.1 X 10 -3 M

15. What mass of solid NaOH (97.0% NaOH by mass) is required to prepare 1.00 L of a 10.0% solution of NaOH by mass? The density of the 10.0% solution is 1.109 g/mL.

ANSWER

114 g

16. Write a balanced molecular equation describing each of the following chemical reactions.
(a) Solid calcium carbonate is heated and decomposes to solid calcium oxide and carbon dioxide gas.
(b) Gaseous butane, C4H10, reacts with diatomic oxygen gas to yield gaseous carbon dioxide and water vapor.
(c) Aqueous solutions of magnesium chloride and sodium hydroxide react to produce solid magnesium hydroxide and aqueous sodium chloride.
(d) Water vapor reacts with sodium metal to produce solid sodium hydroxide and hydrogen gas.

ANSWER

a) CaCO3(s)−>CaO(s)+CO2(g)
b) 2 C 4 H 10 (g)+13O2 (g)−>8CO 2 (g)+10H 2 O(g)
c) MgCl2 (aq)+2NaOH(aq)−>Mg(OH)2 (s)+2NaCl(aq)
d) 2H 2 O(g)+2Na(s)−>2NaOH(s)+H 2(g)

Access for free at openstax.org

Read more

Basic Concepts of Chemistry – I

Chemistry: The Central Science

Chemistry is sometimes referred to as “the central science” due to its interconnectedness with a vast array of other STEM disciplines (STEM stands for areas of study in the science, technology, engineering, and math fields). Chemistry and the language of chemists play vital roles in biology, medicine, materials science, forensics, environmental science, and many other fields . The basic principles of physics are essential for understanding many aspects of chemistry, and there is extensive overlap between many subdisciplines within the two fields, such as chemical physics and nuclear chemistry. Mathematics, computer science, and information theory provide important tools that help us calculate, interpret, describe, and generally make sense of the chemical world. Biology and chemistry converge in biochemistry, which is crucial in  understanding the many complex factors and processes that keep living organisms (such as us) alive. Chemical engineering, materials science, and nanotechnology combine chemical principles and empirical findings to produce useful substances, ranging from gasoline to fabrics to electronics. Agriculture, food science, veterinary science, and brewing and wine making help provide sustenance in the form of food and drink to the world’s population. Medicine, pharmacology, biotechnology, and botany identify and produce substances that help keep us healthy. Environmental science, geology, oceanography, and atmospheric science incorporate many chemical ideas to help us better understand and protect our physical world. Chemical ideas are used to help understand the universe in astronomy and cosmology.

The Scientific Method

Chemistry is a science based on observation and experimentation. Doing chemistry involves attempting to answer questions and explain observations in terms of the laws and theories of chemistry, using procedures that are accepted by the scientific community. There is no single route to answering a question or explaining an observation, but there is an aspect common to every approach: Each uses knowledge based on experiments that can be reproduced to verify the results. Some routes involve a hypothesis, a tentative explanation of

observations that acts as a guide for gathering and checking information. A hypothesis is tested by experimentation, calculation, and/or comparison with the experiments of others and then refined as needed.

Some hypotheses are attempts to explain the behavior that is summarized in laws. The laws of science summarize a vast number of experimental observations, and describe or predict some facet of the natural world. If such a hypothesis turns out to be capable of explaining a large body of experimental data, it can reach the status of a theory. Scientific theories are well-substantiated, comprehensive, testable explanations of particular aspects of nature. Theories are accepted because they provide satisfactory explanations, but they can be modified if new data become available. The path of discovery that leads from question and observation to law or hypothesis to theory, combined with experimental verification of the hypothesis and any necessary modification of the theory, is called the scientific method

Phases and Classification of Matter

Matter is defined as anything that occupies space and has mass, and it is all around us. Solids and liquids are more obviously matter: We can see that they take up space, and their weight tells us that they have mass. Gases are also matter; if gases did not take up space, a balloon would not inflate (increase its volume) when filled with gas. Solids, liquids, and gases are the three states of matter commonly found on earth .

A solid is rigid and possesses a definite shape. A liquid flows and takes the shape of its container, except that it forms a flat or slightly curved upper surface when acted upon by gravity. (In zero gravity, liquids assume a spherical shape.) Both liquid and solid samples have volumes that are very nearly independent of pressure. A gas takes both the shape and volume of its container.

A fourth state of matter, plasma, occurs naturally in the interiors of stars. A plasma is a gaseous state of matter that contains appreciable numbers of electrically charged particles . The presence of these charged particles imparts unique properties to plasmas that justify their classification as a state of matter distinct from gases. In addition to stars, plasmas are found in some other high-temperature environments (both natural and man-made), such as lightning strikes, certain television screens, and specialized analytical instruments used to detect trace amounts of metals

Some samples of matter appear to have properties of solids, liquids, and/or gases at the same time. This can occur when the sample is composed of many small pieces. For example, we can pour sand as if it were a liquid because it is composed of many small grains of solid sand. Matter can also have properties of more than one state when it is a mixture, such as with clouds. Clouds appear to behave somewhat like gases, but they are actually mixtures of air (gas) and tiny particles of water (liquid or solid).

The mass of an object is a measure of the amount of matter in it. One way to measure an object’s mass is to measure the force it takes to accelerate the object. It takes much more force to accelerate a car than a bicycle because the car has much more mass. A more common way to determine the mass of an object is to use a balance to compare its mass with a standard mass.

Although weight is related to mass, it is not the same thing. Weight refers to the force that gravity exerts on an object. This force is directly proportional to the mass of the object. The weight of an object changes as the force of gravity changes, but its mass does not. An astronaut’s mass does not change just because she goes to the moon. But her weight on the moon is only one-sixth her earth-bound weight because the moon’s gravity is only one-sixth that of the earth’s. She may feel “weightless” during her trip when she experiences negligible external forces (gravitational or any other), although she is, of course, never “massless.”

The law of conservation of matter summarizes many scientific observations about matter: It states that there is no detectable change in the total quantity of matter present when matter converts from one type to another (a chemical change) or changes among solid, liquid, or gaseous states (a physical change).

Classifying Matter

Matter can be classified into several categories. Two broad categories are mixtures and pure substances. A pure substance has a constant composition. All specimens of a pure substance have exactly the same makeup and properties. Any sample of sucrose (table sugar) consists of 42.1% carbon, 6.5% hydrogen, and 51.4% oxygen by mass. Any sample of sucrose also has the same physical properties, such as melting point, color, and sweetness, regardless of the source from which it is isolated.

Pure substances may be divided into two classes: elements and compounds. Pure substances that cannot be broken down into simpler substances by chemical changes are called elements. Iron, silver, gold, aluminum, sulfur, oxygen, and copper are familiar examples of the more than 100 known elements, of which about 90 occur naturally on the earth, and two dozen or so have been created in laboratories.

Pure substances that are comprised of two or more elements are called compounds. Compounds may be broken down by chemical changes to yield either elements or other compounds, or both. Mercury(II) oxide, an orange, crystalline solid, can be broken down by heat into the elements mercury and oxygen .

When heated in the absence of air, the compound sucrose is broken down into the element carbon and the compound water. (The initial stage of this process, when the sugar is turning brown, is known as caramelization—this is what imparts the characteristic sweet and nutty flavor to caramel apples, caramelized onions, and caramel). Silver(I) chloride is a white solid that can be broken down into its elements, silver and chlorine, by absorption of light. This property is the basis for the use of this compound in photographic films and photochromic eyeglasses (those with lenses that darken when exposed to light

The properties of combined elements are different from those in the free, or uncombined, state. For example, white crystalline sugar (sucrose) is a compound resulting from the chemical combination of the element carbon, which is a black solid in one of its uncombined forms, and the two elements hydrogen and oxygen, which are colorless gases when uncombined. Free sodium, an element that is a soft, shiny, metallic solid, and free chlorine, an element that is a yellow-green gas, combine to form sodium chloride (table salt), a compound that is a white, crystalline solid.

A mixture is composed of two or more types of matter that can be present in varying amounts and can be separated by physical changes, such as evaporation (you will learn more about this later). A mixture with a composition that varies from point to point is called a heterogeneous mixture. Some examples of heterogeneous mixtures are chocolate chip cookies (we can see the separate bits of chocolate, nuts, and cookie dough) and granite (we can see the quartz, mica, feldspar, and more).

Although there are just over 100 elements, tens of millions of chemical compounds result from different combinations of these elements. Each compound has a specific composition and possesses definite chemical and physical properties that distinguish it from all other compounds. And, of course, there are innumerable ways to combine elements and compounds to form different mixtures. A summary of how to distinguish between the various major classifications of matter is shown in the below figure

Eleven elements make up about 99% of the earth’s crust and atmosphere . Oxygen constitutes nearly one-half and silicon about one-quarter of the total quantity of these elements. A majority of elements on earth are found in chemical combinations with other elements; about one-quarter of the elements are also found in the free state

Elemental Composition of Earth

Element Symbol Percent Mass Element   Symbol   Percent Mass
  oxygen O 49.20 chlorine Cl 0.19
silicon Si 25.67 phosphorus P 0.11
aluminum Al 7.50 manganese Mn 0.09
iron Fe 4.71 carbon C 0.08
calcium Ca 3.39 sulfur S 0.06
sodium Na 2.63 barium Ba 0.04
potassium K 2.40 nitrogen N 0.03
magnesium Mg 1.93 fluorine F 0.03
hydrogen H 0.87 strontium Sr 0.02
titanium Ti 0.58 all others 0.47

Atoms and Molecules

An atom is the smallest particle of an element that has the properties of that element and can enter into a chemical combination. Consider the element gold, for example. Imagine cutting a gold nugget in half, then cutting one of the halves in half, and repeating this process until a piece of gold remained that was so small that it could not be cut in half (regardless of how tiny your knife may be). This minimally sized piece of gold is an atom (from the Greek atomos, meaning “indivisible”). This atom would no longer be gold if it were divided any further

An atom is so small that its size is difficult to imagine. One of the smallest things we can see with our unaided eye is a single thread of a spider web: These strands are about 1/10,000 of a centimeter (0.0001 cm) in diameter. Although the cross-section of one strand is almost impossible to see without a microscope, it is huge on an atomic scale

The above images provide an increasingly closer view: (a) a cotton boll, (b) a single cotton fiber viewed under an optical microscope (magnified 40 times), (c) an image of a cotton fiber obtained with an electron microscope (much higher magnification than with the optical microscope); and (d and e) atomic-level models of the fiber (spheres of different colors represent atoms of different elements). (credit c: modification of work by “Featheredtar”/Wikimedia Commons)

An atom is so light that its mass is also difficult to imagine. A billion lead atoms (1,000,000,000 atoms) weigh about 3 X 10−13 grams, a mass that is far too light to be weighed on even the world’s most sensitive balances. It would require over 300,000,000,000,000 lead atoms (300 trillion, or 3      1014) to be weighed, and they would weigh only 0.0000001 gram

It is rare to find collections of individual atoms. Only a few elements, such as the gases helium, neon, and argon, consist of a collection of individual atoms that move about independently of one another. Other elements, such as the gases hydrogen, nitrogen, oxygen, and chlorine, are composed of units that consist of pairs of atoms . One form of the element phosphorus consists of units composed of four phosphorus atoms. The element sulfur exists in various forms, one of which consists of units composed of eight sulfur atoms. These units are called molecules. A molecule consists of two or more atoms joined by strong forces called chemical bonds. The atoms in a molecule move around as a unit, much like the cans of soda in a six-pack or a bunch of keys joined together on a single key ring. A molecule may consist of two or more identical atoms, as in the molecules found in the elements hydrogen, oxygen, and sulfur, or it may consist of two or more different atoms, as in the molecules found in water. Each water molecule is a unit that contains two hydrogen atoms and one oxygen atom. Each glucose molecule is a unit that contains 6 carbon atoms, 12 hydrogen atoms, and 6 oxygen atoms. Like atoms, molecules are incredibly small and light.

Physical and Chemical Properties

The characteristics that distinguish one substance from another are called properties. A physical property is a characteristic of matter that is not associated with a change in its chemical composition. Familiar examples of physical properties include density, color, hardness, melting and boiling points, and electrical conductivity.

Some physical properties, such as density and color, may be observed without changing the physical state of the matter. Other physical properties, such as the melting temperature of iron or the freezing temperature of water, can only be observed as matter undergoes a physical change. A physical change is a change in the state or properties of matter without any accompanying change in the chemical identities of the substances contained in the matter. Physical changes are observed when wax melts, when sugar dissolves in coffee, and when steam condenses into liquid water . Other examples of physical changes include magnetizing and demagnetizing metals (as is done with common antitheft security tags) and grinding solids into powders (which can sometimes yield noticeable changes in color). In each of these examples, there is a change in the physical state, form, or properties of the substance, but no change in its chemical composition

(a) Wax undergoes a physical change when solid wax is heated and forms liquid wax. (b) Steam condensing inside a cooking pot is a physical change, as water vapor is changed into liquid water. (credit a: modification of work by “95jb14”/Wikimedia Commons; credit b: modification of work by “mjneuby”/Flickr)

The change of one type of matter into another type (or the inability to change) is a chemical property. Examples of chemical properties include flammability, toxicity, acidity, and many other types of reactivity. Iron, for example, combines with oxygen in the presence of water to form rust; chromium does not oxidize . Nitroglycerin is very dangerous because it explodes easily; neon poses almost no hazard because it is very unreactive

(a) One of the chemical properties of iron is that it rusts; (b) one of the chemical properties of chromium is that it does not. (credit a: modification of work by Tony Hisgett; credit b: modification of work by “Atoma”/Wikimedia Commons)

A chemical change always produces one or more types of matter that differ from the matter present before the change. The formation of rust is a chemical change because rust is a different kind of matter than the iron, oxygen, and water present before the rust formed. The explosion of nitroglycerin is a chemical change because the gases produced are very different kinds of matter from the original substance. Other examples of chemical changes include reactions that are performed in a lab (such as copper reacting with nitric acid), all forms of combustion (burning), and food being cooked, digested, or rotting

While many elements differ dramatically in their chemical and physical properties, some elements have similar properties. For example, many elements conduct heat and electricity well, whereas others are poor conductors. These properties can be used to sort the elements into three classes: metals (elements that conduct well), nonmetals (elements that conduct poorly), and metalloids (elements that have intermediate conductivities).

The periodic table is a table of elements that places elements with similar properties close together . You will learn more about the periodic table as you continue your study of chemistry

The periodic table shows how elements may be grouped according to certain similar properties. Note the background color denotes whether an element is a metal, metalloid, or nonmetal, whereas the element symbol color indicates whether it is a solid, liquid, or gas.

Measurements

Measurements provide much of the information that informs the hypotheses, theories, and laws describing the behavior of matter and energy in both the macroscopic and microscopic domains of chemistry. Every measurement provides three kinds of information:

  1. the size or magnitude of the measurement (a number);
  2. a standard of comparison for the measurement (a unit); and
  3.  an indication of the uncertainty of the measurement.

While the number and unit are explicitly represented when a quantity is written, the uncertainty is an aspect of the measurement result that is more implicitly represented and will be discussed later.

The number in the measurement can be represented in different ways, including decimal form and scientific notation. (Scientific notation is also known as exponential notation). For example, the maximum takeoff weight of a Boeing 777-200ER airliner is 298,000 kilograms, which can also be written as 2.98 X 105 kg. The mass of the average mosquito is about 0.0000025 kilograms ,  which can be written as  2.5 X 10-6  Kg  .

Units, such as liters, pounds, and centimeters, are standards of comparison for measurements. A 2-liter   

 bottle of a soft drink contains a volume of beverage that is twice that of the accepted volume of 1 liter.  Without units, a number can be meaningless, confusing, or possibly life threatening. Suppose a doctor prescribes phenobarbital to control a patient’s seizures and states a dosage of “100” without specifying units. Not only will this be confusing to the medical professional giving the dose, but the consequences can be dire.

The measurement units for seven fundamental properties (“base units”) are listed in the below table . The standards for these units are fixed by international agreement, and they are called the International System of Units or SI Units (from the French, Le Système International d’Unités). SI units have been used by the United States National Institute of Standards and Technology (NIST) since 1964. Units for other properties may be derived from these seven base units.

Base Units of the SI System

length meter m
mass kilogram kg
time second s
temperature kelvin K
electric current ampere A
amount of substance mole mol
luminous intensity candela cd

Everyday measurement units are often defined as fractions or multiples of other units . Fractional or multiple  SI units are named using a prefix and the name of the base unit. For example, a length of 1000 meters is also called a kilometer because the prefix kilo means “one thousand,” which in scientific notation is 103 (1 kilometer = 1000 m = 103 m). The prefixes used and the powers to which 10 are raised are listed in  the below table

Common Unit Prefixes                                               

femto f 10−15 1 femtosecond (fs) = 1       10−15 s (0.000000000000001 s)
pico p 10−12 1 picometer (pm) = 1       10−12 m (0.000000000001 m)

nano n 10−9 4 nanograms (ng) = 4    10−9 g (0.000000004 g)
micro µ 10−6 1 microliter (μL) = 1        10−6 L (0.000001 L)
milli m 10−3 2 millimoles (mmol) = 2      10−3 mol (0.002 mol)
centi c 10−2 7 centimeters (cm) = 7     10−2 m (0.07 m)
deci d 10−1 1 deciliter (dL) = 1     10−1 L (0.1 L )
kilo k 103 1 kilometer (km) = 1       103 m (1000 m)
mega M 106 3 megahertz (MHz) = 3     106 Hz (3,000,000 Hz)
giga G 109 8 gigayears (Gyr) = 8     109 yr (8,000,000,000 yr)
tera T 1012 5 terawatts (TW) = 5      1012 W (5,000,000,000,000 W)

SI Base Units

The initial units of the metric system, which eventually evolved into the SI system, were established in France during the French Revolution. The original standards for the meter and the kilogram were adopted there in 1799 and eventually by other countries. This section introduces four of the SI base units commonly used in chemistry. Other SI units will be introduced in subsequent chapters.

Length

The standard unit of length in both the SI and original metric systems is the meter (m). A meter was originally specified as 1/10,000,000 of the distance from the North Pole to the equator. It is now defined as the distance light in a vacuum travels in 1/299,792,458 of a second. A meter is about 3 inches longer than a yard  , one meter is about 39.37 inches or 1.094 yards. Longer distances are often reported in kilometers (1 km

= 1000 m = 103 m), whereas shorter distances can be reported in centimeters (1 cm = 0.01 m = 10−2 m) or

millimeters (1 mm = 0.001 m = 10−3 m).

Above figure gives The relative lengths of 1 m, 1 yd, 1 cm, and 1 in. are shown (not actual size), as well as comparisons of 2.54 cm and  1 in., and of 1 m and 1.094 yd

Mass

The standard unit of mass in the SI system is the kilogram (kg). The kilogram was previously defined by the International Union of Pure and Applied Chemistry (IUPAC) as the mass of a specific reference object. This object was originally one liter of pure water, and more recently it was a metal cylinder made from a platinum- iridium alloy with a height and diameter of 39 mm as shown in below figure . In May 2019, this definition was changed to one that is based instead on precisely measured values of several fundamental physical constants. One kilogram is about 2.2 pounds. The gram (g) is exactly equal to 1/1000 of the mass of the kilogram (10−3 kg)

Temperature

Temperature is an intensive property. The SI unit of temperature is the kelvin (K). The IUPAC convention is to use kelvin (all lowercase) for the word, K (uppercase) for the unit symbol, and neither the word “degree” nor the degree symbol (°). The degree Celsius (°C) is also allowed in the SI system, with both the word “degree” and the degree symbol used for Celsius measurements. Celsius degrees are the same magnitude as those of kelvin, but the two scales place their zeros in different places. Water freezes at 273.15 K (0 °C) and boils at 373.15 K (100 °C) by definition, and normal human body temperature is approximately 310 K (37 °C).

Time

The SI base unit of time is the second (s). Small and large time intervals can be expressed with the appropriate prefixes; for example, 3 microseconds = 0.000003 s = 3×10−6 and 5 megaseconds = 5,000,000 s = 5  106 s. Alternatively, hours, days, and years can be used.

Derived SI Units

We can derive many units from the seven SI base units. For example, we can use the base unit of length to define a unit of volume, and the base units of mass and length to define a unit of density.

Volume

Volume is the measure of the amount of space occupied by an object. The standard SI unit of volume is defined by the base unit of length . The standard volume is a cubic meter (m3), a cube with an edge length of exactly one meter. To dispense a cubic meter of water, we could build a cubic box with edge lengths of exactly one meter. This box would hold a cubic meter of water or any other substance.

A more commonly used unit of volume is derived from the decimeter (0.1 m, or 10 cm). A cube with edge lengths of exactly one decimeter contains a volume of one cubic decimeter (dm3). A liter (L) is the more common name for the cubic decimeter. One liter is about 1.06 quarts.

A cubic centimeter (cm3) is the volume of a cube with an edge length of exactly one centimeter. The abbreviation cc (for cubic centimeter) is often used by health professionals. A cubic centimeter is equivalent to a milliliter (mL) and is 1/1000 of a liter.

The relative volumes are shown for cubes of 1 m3, 1 dm3 (1 L), and 1 cm3 (1 mL) (not to scale). (b) The diameter of a dime is compared relative to the edge length of a 1-cm3 (1-mL) cube.

Density

We use the mass and volume of a substance to determine its density. Thus, the units of density are defined by the base units of mass and length. The density of a substance is the ratio of the mass of a sample of the substance to its volume. The SI unit for density is the kilogram per cubic meter (kg/m3). For many situations, however, this is an inconvenient unit, and we often use grams per cubic centimeter (g/cm3) for the densities of solids and liquids, and grams per liter (g/L) for gases. Although there are exceptions, most liquids and solids have densities that range from about 0.7 g/cm3 (the density of petrol ) to 19 g/cm3 (the density of gold). The density of air is about 1.2 g/L

Densities of Common Substances

Solids Liquids   Gases (at 25 °C and 1 atm)
ice (at 0 °C) 0.92 g/cm3 water 1.0 g/cm3 dry air 1.20 g/L
oak (wood) 0.60–0.90 g/cm3 ethanol 0.79 g/cm3 oxygen 1.31 g/L
iron 7.9 g/cm3 acetone 0.79 g/cm3 nitrogen 1.14 g/L
copper 9.0 g/cm3 glycerin 1.26 g/cm3 carbon dioxide 1.80 g/L
lead 11.3 g/cm3 olive oil 0.92 g/cm3 helium 0.16 g/L
silver 10.5 g/cm3 gasoline 0.70–0.77 g/cm3 neon 0.83 g/L
gold 19.3 g/cm3 mercury 13.6 g/cm3 radon 9.1 g/L

While there are many ways to determine the density of an object, perhaps the most straightforward method involves separately finding the mass and volume of the object, and then dividing the mass of the sample by its volume. In the following example, the mass is found directly by weighing, but the volume is found indirectly through length measurements

Density = Mass / Volume

Measurement Uncertainty, Accuracy, and Precision

Counting is the only type of measurement that is free from uncertainty, provided the number of objects being counted does not change while the counting process is underway. The result of such a counting measurement is an example of an exact number. By counting the eggs in a carton, one can determine exactly how many eggs the carton contains. The numbers of defined quantities are also exact. By definition, 1 foot is exactly 12 inches, 1 inch is exactly 2.54 centimeters, and 1 gram is exactly 0.001 kilogram. Quantities derived from measurements other than counting, however, are uncertain to varying extents due to practical limitations of the measurement process used.

Significant Figures in Measurement

The numbers of measured quantities, unlike defined or directly counted quantities, are not exact. To measure the volume of liquid in a graduated cylinder (below figure) ,  you should make a reading at the bottom of the meniscus, the lowest point on the curved surface of the liquid

Refer to the above  illustration . The bottom of the meniscus in this case clearly lies between the 21 and 22 markings, meaning the liquid volume is certainly greater than 21 mL but less than 22 mL. The meniscus appears to be a bit closer to the 22-mL mark than to the 21-mL mark, and so a reasonable estimate of the liquid’s volume would be 21.6 mL. In the number 21.6, then, the digits 2 and 1 are certain, but the 6 is an estimate. Some people might estimate the meniscus position to be equally distant from each of the markings and estimate the tenth-place digit as 5, while others may think it to be even closer to the 22-mL mark and estimate this digit to be 7. Note that it would be pointless to attempt to estimate a digit for the hundredths place, given that the tenths-place digit is uncertain. In general, numerical scales such as the one on this graduated cylinder will permit measurements to one-tenth of the smallest scale division. The scale in this case has 1-mL divisions, and so volumes may be measured to the nearest 0.1 mL.

This concept holds true for all measurements, even if you do not actively make an estimate. If you place a quarter on a standard electronic balance, you may obtain a reading of 6.72 g. The digits 6 and 7 are certain, and the 2 indicates that the mass of the quarter is likely between 6.71 and 6.73 grams. The quarter weighs about 6.72 grams, with a nominal uncertainty in the measurement of ± 0.01 gram. If the coin is weighed on a more sensitive balance, the mass might be 6.723 g. This means its mass lies between 6.722 and 6.724 grams, an uncertainty of 0.001 gram. Every measurement has some uncertainty, which depends on the device used (and the user’s ability). All of the digits in a measurement, including the uncertain last digit, are called significant figures or significant digits. Note that zero may be a measured value; for example, if you stand on a scale that shows weight to the nearest Kg ,  and it shows “120,” then the 1 (hundreds), 2 (tens) and 0 (ones) are all significant (measured) values.

Significant Figures in Calculations

A second important principle of uncertainty is that results calculated from a measurement are at least as uncertain as the measurement itself. Take the uncertainty in measurements into account to avoid misrepresenting the uncertainty in calculated results. One way to do this is to report the result of a calculation with the correct number of significant figures, which is determined by the following three rules for rounding numbers:

  1. When adding or subtracting numbers, round the result to the same number of decimal places as the         number with the least number of decimal places (the least certain value in terms of addition and subtraction).
  2. When multiplying or dividing numbers, round the result to the same number of digits as the number with the least number of significant figures (the least certain value in terms of multiplication and division).
  3. If the digit to be dropped (the one immediately to the right of the digit to be retained) is less than 5, “round down” and leave the retained digit unchanged; if it is more than 5, “round up” and increase the retained digit by 1. If the dropped digit is 5, and it’s either the last digit in the number or it’s followed only by zeros, round up or down, whichever yields an even value for the retained digit. If any nonzero digits follow the dropped 5, round up. (The last part of this rule may strike you as a bit odd, but it’s based on reliable statistics and is aimed at avoiding any bias when dropping the digit “5,” since it is equally close to both possible values of the retained digit.)

Conversion Factors and Dimensional Analysis

A ratio of two equivalent quantities expressed with different measurement units can be used as a unit conversion factor. For example, the lengths of 2.54 cm and 1 inch. are equivalent (by definition), and so a unit conversion factor may be derived from the ratio

Common Conversion Factors

Length    Volume       Mass
1 m = 1.0936 yd 1 L = 1.0567 qt 1 kg = 2.2046 lb
1 in. = 2.54 cm (exact) 1 qt = 0.94635 L 1 lb = 453.59 g
1 km = 0.62137 mi 1 ft3 = 28.317 L 1 (avoirdupois) oz = 28.349 g
1 mi = 1609.3 m 1 tbsp = 14.787 mL 1 (troy) oz = 31.103 g

When a quantity (such as distance in inches) is multiplied by an appropriate unit conversion factor, the quantity is converted to an equivalent value with different units (such as distance in centimeters).

Take Quiz

1. Classify each of the following as an element, a compound, or a mixture:

  • sucrose
  • copper
  • water
  • nitrogen
  • sulfur
  • air

ANSWER

Copper – Element
Water – Compound
Nitrogen – Element
Sulfur – Element
Air – Mixture
Sucrose – Compound

2. Classify each of the following changes as physical or chemical:
a) condensation of steam
b) burning of gasoline
c) souring of milk
d) dissolving of sugar in water
e) melting of gold

ANSWER

a) Condensation of steam – Physical change
b) Burning of gasoline – Chemical change
c) Souring of milk – Chemical change
d) Dissolving of sugar in water – Physical change
e) Melting of gold – Physical change

3. Indicate the SI base units or derived units that are appropriate for the following measurements:

  • the volume of a flu shot or a measles vaccination
  • the mass of the moon
  • the distance from Mumbai to Kolkata City
  • the speed of sound
  • the density of air

ANSWER

The mass of the moon: Kilogram (kg)
The distance from Mumbai to Kolkata City: Meter (m)
The speed of sound: Meters per second (m/s)
The density of air: Kilograms per cubic meter (kg/m³)
The volume of a flu shot or a measles vaccination: Cubic meter (m³) or more commonly, milliliter (mL) for practical use in this context

4. Give the name and symbol of the prefixes used with SI units to indicate multiplication by the following exact quantities.

(a) 103
(b) 10-2
c) 0.1
d) 10-3
(e) 1,000,000
(f ) 0.000001

ANSWER

(a) 103
– Name: Kilo
– Symbol: k
(b) 10-2
– Name: Centi
– Symbol: c
(c) 0.1
– Name: Deci
– Symbol: d
(d) 10-3
– Name: Milli
– Symbol: m
(e) 1,000,000
– Name: Mega
– Symbol: M
(f) 0.000001
– Name: Micro
– Symbol: µ

5. Give the name of the prefix and the quantity indicated by the following symbols that are used with SI base units.
1. m
2. c
3. d
4. G
5. k

ANSWER

c
– Name: Centi
– Quantity: 10−210^{-2}10−2 or 0.01
d
– Name: Deci
– Quantity: 10−110^{-1}10−1 or 0.1
G
– Name: Giga
– Quantity: 10910^9109 or 1,000,000,000

k
– Name: Kilo
– Quantity: 10310^3103 or 1,000

m
– Name: Milli
– Quantity: 10−310^{-3}10−3 or 0.001

6. When elemental iron corrodes it combines with oxygen in the air to ultimately form red brown iron(III) oxide called rust. (a) If a shiny iron nail with an initial mass of 23.2 g is weighed after being coated in a layer of rust, would you expect the mass to have increased, decreased, or remained the same?

ANSWER

When elemental iron corrodes and combines with oxygen from the air to form iron(III) oxide (rust), the mass of the iron nail would increase. This is because the process of rusting involves the addition of oxygen atoms to the iron atoms, resulting in an increase in the overall mass of the nail

7. Yeast converts glucose to ethanol and carbon dioxide during anaerobic fermentation as depicted in the simple chemical equation here
Glucose ———— Ethanol + Carbondioxide
If 200.0 g of glucose is fully converted, what will be the total mass of ethanol and carbon dioxide produced ?

ANSWER

The law of conservation of mass states that the mass of the reactants in a chemical reaction is equal to the mass of the products. Therefore, if 200.0 g of glucose is fully converted during anaerobic fermentation, the total mass of ethanol and carbon dioxide produced will also be 200.0 g

8. The volume of a sample of oxygen gas changed from 10 mL to 11 mL as the temperature changed. Is this a chemical or physical change?

ANSWER

The change in the volume of a sample of oxygen gas from 10 mL to 11 mL as the temperature changes is a physical change. This is because the change involves a physical property (volume) and is reversible, without altering the chemical composition or identity of the oxygen gas.

9. A 2.0-liter volume of hydrogen gas combined with 1.0 liter of oxygen gas to produce 2.0 liters of water vapor. Does oxygen undergo a chemical or physical change?

ANSWER

In the reaction where 2.0 liters of hydrogen gas combine with 1.0 liter of oxygen gas to produce 2.0 liters of water vapor, oxygen undergoes a chemical change. This is because the oxygen molecules react with hydrogen molecules to form a new substance, water (H₂O), which has different chemical properties and composition from the reactant gases.

10. Indicate the SI base units or derived units that are appropriate for the following measurements:
1. the length of a marathon race (26 miles 385 yards)
2. the mass of an automobile
3. the volume of a swimming pool
4. the speed of an airplane
5. the density of gold
6. the area of a football field

ANSWER

Here are the appropriate SI base units or derived units for each measurement:
1. The length of a marathon race (26 miles 385 yards)
– Meter (m)
2. The mass of an automobile
– Kilogram (kg)
3. The volume of a swimming pool
– Cubic meter (m³)
4. The speed of an airplane
– Meters per second (m/s)
5. The density of gold
– Kilograms per cubic meter (kg/m³)
6. The area of a football field
– Square meter (m²)

Access for free at openstax.org

Read more

GRAVITATION

What do a falling apple and the orbit of the moon have in common? You will learn in this chapter that each is caused by gravitational force. The motion of all celestial objects, in fact, is determined by the gravitational force, which depends on their mass and separation.

Johannes Kepler discovered three laws of planetary motion that all orbiting planets and moons follow. Years later, Isaac Newton found these laws useful in developing his law of universal gravitation. This law relates gravitational force to the masses of objects and the distance between them.

Kepler’s Laws of Planetary Motion

Hundreds of artificial satellites orbit Earth together with thousands of pieces of debris. The moon’s orbit around Earth has intrigued humans from time immemorial. The orbits of planets, asteroids, meteors, and comets around the sun are no less interesting. If we look farther, we see almost unimaginable numbers of stars, galaxies, and other celestial objects orbiting one another and interacting through gravity.

All these motions are governed by gravitational force. The orbital motions of objects in our own solar system are simple enough to describe with a few fairly simple laws. The orbits of planets and moons satisfy the following two conditions:

  • The mass of the orbiting object, m, is small compared to the mass of the object it orbits, M.
  • The system is isolated from other massive objects.

Based on the motion of the planets about the sun, Kepler devised a set of three classical laws, called Kepler’s laws of planetary motion, that describe the orbits of all bodies satisfying these two conditions:

  1. The orbit of each planet around the sun is an ellipse with the sun at one focus.
  2. Each planet moves so that an imaginary line drawn from the sun to the planet sweeps out equal areas in equal times.
  3. The ratio of the squares of the periods of any two planets about the sun is equal to the ratio of the cubes of their average distances from the sun.

Kepler’s First Law

The orbit of each planet about the sun is an ellipse with the sun at one focus, as shown in the below figure . The planet’s closest approach to the sun is called aphelion and its farthest distance from the sun is called perihelion

a) An ellipse is a closed curve such that the sum of the distances from a point on the curve to the two foci (f1 and f2) is constant.

(b) For any closed orbit, m follows an elliptical path with M at one focus.

(c) The aphelion (ra) is the closest distance between the planet and the sun, while the perihelion (rp) is the farthest distance from the sun

Hence , If you know the aphelion (ra) and perihelion (rp) distances, then you can calculate the semi-major axis (a) and semi-minor axis (b).

a = (ra + rp)/2

b = √rarp

Kepler’s Second Law

Each planet moves so that an imaginary line drawn from the sun to the planet sweeps out equal areas in equal times, as shown in the below figure

The shaded regions have equal areas. The time for m to go from A to B is the same as the time to go from C to D and from E to F. The mass m moves fastest when it is closest to M. Kepler’s second law was originally devised for planets orbiting the sun, but it has broader validity.

Kepler’s Third Law

The ratio of the periods squared of any two planets around the sun is equal to the ratio of their average distances from the sun cubed. In equation form, this is

where T is the period (time for one orbit) and r is the average distance (also called orbital radius). This equation is valid only for comparing two small masses orbiting a single large mass. Most importantly, this is only a descriptive equation; it gives no information about the cause of the equality.

Newton’s Universal Law of Gravitation :

Newton was the first to propose an exact mathematical form and to use that form to show that the motion of heavenly bodies should be conic sections—circles, ellipses, parabolas, and hyperbolas. This theoretical prediction was a major triumph. It had been known for some time that moons, planets, and comets follow such paths, but no one had been able to propose an explanation of the mechanism that caused them to follow these paths and not others.

The gravitational force between two bodies  is always attractive and depends on the masses involved and the distance between them. Expressed in modern language, Newton’s universal law of gravitation states that every object in the universe attracts every other object with a force that is directed along a line joining them. The force is directly proportional to the product of their masses and inversely proportional to the square of the distance between them. This attraction is illustrated by the below figure

Gravitational attraction is along a line joining the centers of mass (CM) of the two bodies. The magnitude of the force on each body is the same, consistent with Newton’s third law (action-reaction).

For two bodies having masses m and M with a distance r between their centers of mass, the equation for Newton’s universal law of gravitation is

where F is the magnitude of the gravitational force and G is a proportionality factor called the gravitational constant. G is a universal constant, meaning that it is thought to be the same everywhere in the universe. It has been measured experimentally to be

If a person has a mass of 60.0 kg, what would be the force of gravitational attraction on him at Earth’s surface? G is given above, Earth’s mass M is 5.97 × 1024 kg, and the radius r of Earth is 6.38 × 106 m. Putting these values into Newton’s universal law of gravitation gives

We can check this result with the relationship:

You may remember that g, the acceleration due to gravity, is another important constant related to gravity. By substituting g for a in the equation for Newton’s second law of motion we get F = mg. Combining this with the equation for universal gravitation gives

Cancelling the mass m on both sides of the equation and filling in the values for the gravitational constant and mass and radius of the Earth, gives the value of g, which may look familiar.

This is a good point to recall the difference between mass and weight. Mass is the amount of matter in an object; weight is the force of attraction between the mass within two objects. Weight can change because g is different on every moon and planet. An object’s mass m does not change but its weight mg can.

We can also derive Kepler’s third law from Newtons’s Universal Law of Gravitation . Applying Newton’s second law of motion to angular motion gives an expression for centripetal force, which can be equated to the expression for force in the universal gravitation equation. This expression can be manipulated to produce the equation for Kepler’s third law. We saw earlier that the expression r3/T2 is a constant for satellites orbiting the same massive object. The derivation of Kepler’s third law from Newton’s law of universal gravitation and Newton’s second law of motion yields that constant:

               r 3 / T2 =    GM / 4π2

where M is the mass of the central body about which the satellites orbit (for example, the sun in our solar system).

The universal gravitational constant G is determined experimentally. This definition was first done accurately in 1798 by English scientist Henry Cavendish (1731–1810), more than 100 years after Newton published his universal law of gravitation. The measurement of G is very basic and important because it determines the strength of one of the four forces in nature.

Cavendish’s experiment was very difficult because he measured the tiny gravitational attraction between two ordinary-sized masses (tens of kilograms at most) by using an apparatus like that in the below figure .  Remarkably, his value for G differs by less than 1% from the modern value.

Cavendish used an apparatus like this to measure the gravitational attraction between two suspended spheres (m) and two spheres on a stand (M) by observing the amount of torsion (twisting) created in the fiber. The distance between the masses can be varied to check the dependence of the force on distance. Modern experiments of this type continue to explore gravity

Gravitational Potential Energy

 Gravitational potential energy is the stored energy an object has as a result of its position above Earth’s surface (or another object in space). A roller coaster car at the top of a hill has gravitational potential energy.

Let’s examine how doing work on an object changes the object’s energy. If we apply force to lift a rock off the ground, we increase the rock’s potential energy, PE. If we drop the rock, the force of gravity increases the rock’s kinetic energy as the rock moves downward until it hits the ground.

The force we exert to lift the rock is equal to its weight, w, which is equal to its mass, m, multiplied by acceleration due to gravity, g.

f = w = mg

The work we do on the rock equals the force we exert multiplied by the distance, d, that we lift the rock. The work we do on the rock also equals the rock’s gain in gravitational potential energy, PEe.

W = PEe = fmg

Potential energy is particularly useful for forces that change with position, as the gravitational force does over large distances. The change in gravitational potential energy near Earth’s surface is ∆U = mg(y2 – y1). This works very well if g does not change significantly between y1and y2. We return to the definition of work and potential energy to derive an expression that is correct over larger distances

We know that work (W) is the integral of the dot product between force and distance. Essentially, it is the product of the component of a force along a displacement times that displacement. We define ∆U as the negative of the work done by the force we associate with the potential energy. For clarity, we derive an expression for moving a mass m from distance r1 from the center of Earth to distance r2. However, the result can easily be generalized to any two objects changing their separation from one value to another.

Consider the below figure , in which we take m from a distance r1 from Earth’s center to a distance that is r2 from the center. Gravity is a conservative force (its magnitude and direction are functions of location only), so we can take any path we wish, and the result for the calculation of work is the same. We take the path shown, as it greatly simplifies the integration. We first move radially outward from distance r1 to distance r2 , and then move along the arc of a circle until we reach the final position.

The work integral, which determines the change in potential energy, can be evaluated along the path shown in red.

Note two important items with this definition. First,

The potential energy is zero when the two masses are infinitely far apart. Only the difference in U is important, so the choice of

is merely one of convenience. (Recall that in earlier gravity problems, you were free to take U = 0 at the top or bottom of a building, or anywhere.) Second, note that U becomes increasingly more negative as the masses get closer. That is consistent with what you learned about potential energy in Potential Energy and Conservation of Energy. As the two masses are separated, positive work must be done against the force of gravity, and hence, U increases (becomes less negative). All masses naturally fall together under the influence of gravity, falling from a higher to a lower potential energy.

 Escape velocity

Escape velocity is  defined as the minimum initial velocity of an object that is required to escape the surface of a planet (or any large body like a moon) and never return. As usual, we assume no energy lost to an atmosphere, should there be any.

Consider the case where an object is launched from the surface of a planet with an initial velocity directed away from the planet. With the minimum velocity needed to escape, the object would just come to rest infinitely far away, that is, the object gives up the last of its kinetic energy just as it reaches infinity, where the force of gravity becomes zero.

Since U → 0 as r → ∞, this means the total energy is zero. Thus, we find the escape velocity from the surface of an astronomical body of mass M and radius R by setting the total energy equal to zero. At the surface of the body, the object is located at r1=R  and it has escape velocity V1= V𝑒𝑠𝑐. It reaches r2=∞ with velocity v2=0

The escape velocity is the same for all objects, regardless of mass. Also, we are not restricted to the surface of the planet; R can be any starting point beyond the surface of the planet.

In Potential Energy and Conservation of Energy, we described how to apply conservation of energy for systems with conservative forces. We were able to solve many problems, particularly those involving gravity, more simply using conservation of energy. Those principles and problem-solving strategies apply equally well here. The only change is to place the new expression for potential energy into the conservation of energy equation,

Etot=K1+U1=K2+U2

Satellite Orbits and Energy

The Moon orbits Earth. In turn, Earth and the other planets orbit the Sun. The space directly above our atmosphere is filled with artificial satellites in orbit. We examine the simplest of these orbits, the circular orbit, to understand the relationship between the speed and period of planets and satellites in relation to their positions and the bodies that they orbit.

Nicolaus Copernicus first suggested that Earth and all other planets orbit the Sun in circles. He further noted that orbital periods increased with distance from the Sun. Later analysis by Kepler showed that these orbits are actually ellipses, but the orbits of most planets in the solar system are nearly circular. Earth’s orbital distance from the Sun varies a mere 2%. The exception is the eccentric orbit of Mercury, whose orbital distance varies nearly 40%.

Determining the orbital speed and orbital period of a satellite is much easier for circular orbits, so we make that assumption in the derivation that follows. As we described in the previous section, an object with negative total energy is gravitationally bound and therefore is in orbit. Our computation for the special case of circular orbits will confirm this. We focus on objects orbiting Earth, but our results can be generalized for other cases.

Consider a satellite of mass m in a circular orbit about Earth at distance r from the center of Earth (as shown in the below figure) . It has centripetal acceleration directed toward the center of Earth. Earth’s gravity is the only force acting, so Newton’s second law gives

(A satellite of mass m orbiting at radius r from the center of Earth. The gravitational force supplies the centripetal acceleration.)

We solve for the speed of the orbit, noting that m cancels, to get the orbital speed

We see in the next section that this represents Kepler’s third law for the case of circular orbits. It also confirms Copernicus’s observation that the period of a planet increases with increasing distance from the Sun.

Astronauts in orbit appear to be weightless, as if they were free-falling towards Earth. In fact, they are in free fall. Consider the trajectories shown in the below figure. (This figure is based on a drawing by Newton in his Principia and also appeared earlier in Motion in Two and Three Dimensions.) All the trajectories shown that hit the surface of Earth have less than orbital velocity. The astronauts would accelerate toward Earth along the noncircular paths shown and feel weightless. (Astronauts actually train for life in orbit by riding in airplanes that free fall for 30 seconds at a time.) But with the correct orbital velocity, Earth’s surface curves away from them at exactly the same rate as they fall toward Earth. Of course, staying the same distance from the surface is the point of a circular orbit

(A circular orbit is the result of choosing a tangential velocity such that Earth’s surface curves away at the same rate as the object falls toward Earth)

Energy in Circular Orbits

In Gravitational Potential Energy and Total Energy, we argued that objects are gravitationally bound if their total energy is negative. The argument was based on the simple case where the velocity was directly away or toward the planet. We now examine the total energy for a circular orbit and show that indeed, the total energy is negative. As we did earlier, we start with Newton’s second law applied to a circular orbit,

In the last step, we multiplied by r on each side. The right side is just twice the kinetic energy, so we have

K = ½ mv2  = GmME / 2r

The total energy is the sum of the kinetic and potential energies, so our final result is

E = K + U

   = GmME / 2r  – GmME / r

   = – GmME / 2r

We can see that the total energy is negative, with the same magnitude as the kinetic energy. For circular orbits, the magnitude of the kinetic energy is exactly one-half the magnitude of the potential energy. Remarkably, this result applies to any two masses in circular orbits about their common center of mass, at a distance r from each other. The proof of this is left as an exercise. We will see in the next section that a very similar expression applies in the case of elliptical orbits

We can see that the total energy is negative, with the same magnitude as the kinetic energy. For circular orbits, the magnitude of the kinetic energy is exactly one-half the magnitude of the potential energy. Remarkably, this result applies to any two masses in circular orbits about their common center of mass, at a distance r from each other. The proof of this is left as an exercise. We will see in the next section that a very similar expression applies in the case of elliptical orbits

“Access for free at openstax.org.”

Take Quiz

1) Why do some objects fall faster than others near the surface of the earth if all mass is attracted equally by the force of gravity?
a) Some objects fall faster because of air resistance, which acts in the direction of the motion of the object and exerts more force on objects with less surface area.
b) Some objects fall faster because of air resistance, which acts in the direction opposite the motion of the object and exerts more force on objects with less surface area.
c) Some objects fall faster because of air resistance, which acts in the direction of motion of the object and exerts more force on objects with more surface area.
d) Some objects fall faster because of air resistance, which acts in the direction opposite the motion of the object and exerts more force on objects with more surface area.

ANSWER

b) Some objects fall faster because of air resistance, which acts in the direction opposite the motion of the object and exerts more force on objects with less surface area.

2) Would you have the same mass on the moon as you do on Earth? Would you have the same weight?
a) You would weigh more on the moon than on Earth because gravity on the moon is stronger than gravity on Earth.
b) You would weigh less on the moon than on Earth because gravity on the moon is weaker than gravity on Earth.
c) You would weigh less on the moon than on Earth because gravity on the moon is stronger than gravity on Earth.
d) You would weigh more on the moon than on Earth because gravity on the moon is weaker than gravity on Earth.

ANSWER

b) You would weigh less on the moon than on Earth because gravity on the moon is weaker than gravity on Earth.

3) Calculate Earth’s mass given that the acceleration due to gravity at the North Pole is 9.83 m/s2 and the radius of the Earth is 6371 Km from pole to center.
a) 5.94 X 1017 Kg
b) 5.94 X 1024 Kg
c) 9.36 X 1017 Kg
d) 9.36 X 1024 Kg

ANSWER

b) 5.94 X 1024 Kg

4) Comets have very elongated elliptical orbits with the sun at one focus. Using Kepler’s Law, explain why a comet travels much faster near the sun than it does at the other end of the orbit.
a) Because the satellite sweeps out equal areas in equal times
b) Because the satellite sweeps out unequal areas in equal times
c) Because the satellite is at the other focus of the ellipse
d) Because the square of the period of the satellite is proportional to the cube of its average distance from the sun

ANSWER

a) Because the satellite sweeps out equal areas in equal times

5) Titan, with a radius of 2.58 X 106 m , is the largest moon of the planet Saturn. If the mass of Titan is 1.35 X 1023 Kg , what is the acceleration due to gravity on the surface of this moon?
a) 1.35 m/s2
b) 3.49 m/s2
c) 3.49 X 106 m/s2
d) 1.35 X 106 m/s2

ANSWER

a) 1.35 m/s2

6) A moon orbits a planet in an elliptical orbit. The foci of the ellipse are 50, 000 km apart. The closest approach of the moon to the planet is 400, 000 km. What is the length of the major axis of the orbit?
a) 400, 000 km
b) 450, 000, km
c) 800, 000 km
d) 850, 000 km

ANSWER

d) 850, 000 km

7) The closest Earth comes to the sun is 1.47×108 km, and Earth’s farthest distance from the sun is 1.52×108 km. What is the area inside Earth’s orbit?
a) 2.23×1016 km2
b) 6.79×1016 km2
c) 7.02×1016 km2
d) 7.26×1016 km2

ANSWER

c) 7.02×1016 km2

8) An artificial satellite orbits the Earth at a distance of 1.45×l04 km from Earth’s center. The moon orbits the Earth at a distance of 3.84×105 km once every 27.3 days. How long does it take the satellite to orbit the Earth?
a) 0.200 days
b) 3.07 days
c) 243 days
d) 3721 days

ANSWER

a) 0.200 days

9) Earth is 1.496×108 km from the sun, and Venus is 1.08×108 km from the sun. One day on Venus is 243 Earth days long. What best represents the number of Venusian days in a Venusian year?
a) 0.78 days
b) 0.92 days
c) 1.08 days
d) 1.21 days

ANSWER

b) 0.92 days

10) What is the gravitational force between two 60 Kg people sitting 100m apart ?
a) 2.4 X 10 – 11 N
b) 2.4 X 10 – 9 N
c) 3.6 X 10-1 N
d) 3.6 X 101 N

ANSWER

b) 2.4 X 10 – 9 N

11) The masses of Earth and the moon are 5.97×1024 kg and 7.35×l022 kg, respectively. The distance from Earth to the moon is 3.80×105 km. At what point between the Earth and the moon are the opposing gravitational forces equal? (Use subscripts e and m to represent Earth and moon.)

a) 3.42×105 km from the center of Earth
b) 3.80×105 km from the center of Earth
c) 3.42×106 km from the center of Earth
d) 3.10×107 km from the center of Earth

ANSWER

b) 3.80×105 km from the center of Earth

Read more

Morphology of flowering plants – Part II

Structure of Flower

A complete flower typically has four verticils or whorls. They are as follows –

  1. Calyx – The members that make up this whorl are called sepals. Sepals are green leaf-like structures that protect a flower in its early bud stage. Also, it is the outermost whorl of a flower.
  2. Corolla – The members that make up this whorl are called petals. These structures are brightly coloured and aid in pollination by attracting insects. Also, they can be seen in different shapes like tubular, wheel-shaped, funnel-shaped or bell-shaped. The corolla and calyx together make up a non-reproductive structure called the perianth.
  3. Androecium – They are the male reproductive structures composed of stamens. Each stamen consists of a filament or stalk and an anther. Each anther has pollen sacs that aid in the production of pollen grains.
  4. Gynoecium – They are female reproductive structures composed of carpels. A carpel is composed of 3 parts namely ovary, style and stigma

Androecium: 

It is a male reproductive part. It consists of stamens. Each stamen is made up of filament and anthers.

  • The stamens may either be free (polyandrous) or remain fused together. Epipetalous- stamens are attached to petals.

The cohesion of stamens fused conditions are as follows.

  • Monadelphous: Stamens are united into a single group by union of filaments. The anthers remain free.

 (e.g. Hibiscus)

  • Diadelphous: Stamens are united, and arranged in two bundle s.  (e.g. Pisum)

  • Polyadelphous: Stamens are united and grouped onto several bundles. (e.g. Citrus)

  • Synandrous: Stamens are united their whole length. Both filaments and anthers will be fused. (e.g. Cucurbita)

  • Syngenesious: Stamens are united by the joining of anthers. The filaments remain free. (e.g. Tridax)

The stamens can also be found fused with other whorls. This condition is called adhesion. It may be of different types.

1. Epipetalous: Stamens are found attached (wholly or partially) to the corolla, by means of their filaments. Anthers would be free. (e.g. Ixora)

2. Epiphyllous: When stamens are found attached to the perianth of a flower, it is called epiphyllous condition. (e.g. Gloriosa)

3. Gyandrous: Stamens are attached to the carpel (either by whole length, or by the anthers.  (e.g. Calotropis)

Gynoecium :

The gynoecium of a flower is the female reproductive structure that is most essential for sexual reproduction. The gynoecium has ovaries that produce ovules which eventually develop into seeds and fruit.  Carpel is the main component of gynoecium that consists of ovary, style and stigma.

  1. Stigma is a sticky portion at the distal end of a tubular structure called style. It is the primary receptive site where the pollen grains are captured.
  2. Style is the major structure that aids in the process of fertilization. The pollen tube reaches the ovary via this tubular style.
  3. Ovary is located at the swollen base of the carpel. Style is the elongated tube that links the ovary and the stigma. The ovary produces ovules which are present within chambers known as locules. Based on the number of locules, the ovary can also be classified as multilocular, unilocular, bilocular, etc.
  4. pistil : Stigma, ovary and style are together termed as the pistil. A pistil is considered to be a collection of one or more carpels.

Placentation:

It is defined as the arrangement and attachment of ovules within the plant’s ovary.

The part of the ovary where ovules are attached is the placenta.

Types of placentation:

  • Axile placentation: When the placenta is placed axially as in a lemon, then such type of placentation is termed axile placentation.

  • Marginal placentation: When the placenta forms a ridge along the ventral suture as in pea, then this is known as marginal placentation.

  • Free central placentation: When the ovules are born on the central axis, it is called free central as in primrose and Dianthus.

  • Basal placentation: In this type of placentation, a single ovule is attached to the placenta and placed at the base of the ovary. Eg. Marigold.

Parietal placentation: In this type of placentation, the ovules develop on the inner wall or periphery of the ovary. Eg. Sunflower, kiwi, cucumber, cantaloupe.

Superficial: The placenta is formed all around the inner surface of the ovary including the septa.

Nymphaea (waterlily).

  1. Marginal- e.g. pea
  2. Axile- e.g. lemon, china rose
  3. Parietal- e.g. Argemone, mustard
  4. Free central- e.g. Primrose, Dianthus
  5. Basal- e.g. marigold, sunflower

The Flower

A flower has four whorls; calyx, corolla, androecium and gynoecium. These are attached to the swollen terminal of pedicel called the thalamus

Flower symmetry:

  1. Actinomorphic- radially symmetrical flowers, e.g. chilli, datura, mustard
    1. Zygomorphic- when a flower can be divided into two equal parts in only one vertical plane, e.g. Cassia, pea, etc.

Flowers can be trimerous, tetramerous or pentamerous depending on the multiple of floral appendages present 3, 4 or 5

Types of flowers depending on the presence or absence of bracts (reduced leaf present at the base of pedicel); Bracteate or Ebracteate

Types of flowers based on the position of the ovary:

  1. Hypogynous- gynoecium occupies the highest place, above all the other parts. The ovary is known as superior. The sepals, petals and stamens arise from below the ovary.This is the commonest type of ovary. The flower is technically Hypogynous where hypo means flower parts start ‘below’ the ovary. e.g. brinjal, china rose, mustard

  • Perigynous- gynoecium is present at the same level as the rest of the parts of a flower. The ovary is known as half inferior, The flower parts arise from the rim of a hypanthium – a cup shaped extension of the base which encloses the ovary. Imagine the hypanthium to be a flat ring and you can see that the flower parts actually arise from below the ovary. The flower is technically Perigynous where peri means ‘around’.  e.g. peach, plum, rose

  • Epigynous- thalamus encloses the ovary completely and other parts are present above it. The ovary is known as inferior.  Inferior ovary. The sepals, corolla tube and stamens (filaments plus anthers) arise from above the ovary. The flower is technically Epigynous where epi means ‘above’. e.g. ray florets of a sunflower, guava, cucumber

Aestivation:

The mode of arrangement of sepals or petals in floral bud with respect to the other members of the same whorl is known as aestivation.

The main types of aestivations are valvate, twisted, imbricate and vexillary.

Valvate:

When sepals or petals in a whorl just touch one another at the margin, without overlapping, as in Calotropis, it is said to be valvate.

Twisted:

If one margin of the appendage overlaps that of the next one and so on as in china rose, lady’s finger and cotton, it is called twisted.

Imbricate:

If the margins of sepals or petals overlap one another but not in any particular direction as in Cassia and gulmohur, the aestivation is called imbricate.

Vexillary:

In pea and bean flowers, there are five petals, the largest (standard) overlaps the two lateral petals (wings) which in turn overlap the two smallest anterior petals (keel); this type of aestivation is known as vexillary or papilionaceous.

Families of flowering plants:

Main Difference between Fabaceae, Solanaceae, and Liliaceae

  1. Fabaceae is a legume family, Solanaceae is a potato family, and Liliaceae is the lily family.
  2. Fabaceae contains peas, chickpeas, and soybean, Solanaceae contain potatoes, tomatoes, and bell pepper, while Liliaceae contain onion, lily, and tulip.
  3. Fabaceae is perennials or annuals, Solanaceae is annuals or biennials, while Liliaceae is perennials.
  4. Fabaceae is monocarpellary; Solanaceae is bicarpellary, while Liliaceae is tricarpellary.
  5. Fabaceae contain pods, while Solanaceae and Liliaceae contain berries or capsules.

Refer Flower formula library notes for more information

Fabaceae Solanaceae Liliaceae
It is a large family of trees, shrubs, vines, and herbs. It is a nightshade or potato family of flowering plants These are monocotyledonous, perennial, and herbaceous plants.
Legume family Potato family Lily family
Dicot Dicot Monocot
Peas, chickpeas, and soybean Potato, tomato, and bell pepper. Onion, lily, and tulip
Exhibits marginal placentation. Exhibits axial placentation. Exhibits axial placentation.
Perennial or annuals Perennials, biennials, and annuals. Perennials
Trees, shrubs, vines, and herbs Herbs, shrubs, small trees Herbs and shrubs
Monocarpellary Bicarpellary Tricarpellary
Contain pods Contain berry or capsule Contain berry or capsule
Fix atmospheric nitrogen to improve soil fertility Contain plant varieties with agricultural importance Contain decorative and ornamental plants.

Take Quiz

1. Diadelphous stamens are found in:
a. China rose
b. Citrus
c. Pea
d. China rose and citrus

ANSWER

Diadelphous stamens are found in pea (Fabaceae).

Diadelphous is a condition when stamens are united partially and are present in two brunches, like in Peas. In this condition, filaments of nine separate stamens are fused together forming one unit, and the tenth posterior stamen remains as a stand portion out of the bundle.

2. The sepals, corolla tube and stamens (filaments plus anthers) arise from above the ovary. Based on the position of the ovary it is
a. Hypogyneous
b. Epigyneous
c. Perigyneous

ANSWER

3. The given figure represents a type of aestivation.

The type of aestivation shown in the given figure is known as
a) Valvate
b) Twisted
c) Imbricate
d) Vexillary

ANSWER

4. Axile placentation occurs in
a) Brassicaceae and Solanaceae
b) Asteraceae and Fabaceae
c) Solanaceae and Liliaceae
d) Fabaceae and Liliaceae

ANSWER

5. Gynoecium of Solanaceae is
a) Bicarpellary apocarpous
b) monocarpellary
c) polycarpellary syncarpus
d) bicarpellary syncarpus

ANSWER

In Solanaceae, the gynoecium is bicarpellary, syncarpous, ovary superior, obliquely placed, bilocular, becomes multilocular due to the formation of the false septum, axile placentation is seen.

6. What is the corolla aestivation of the papilionaceous flower which has a large vexillum covering two wings and its keel covered by another wing?
a. Twisted
b. Ascending imbricate
c. Valvate
d. Descending imbricate

ANSWER

Explanation: The corolla of the papilionaceae family has the shape of a butterfly. The butterfly shape is due to the varying size of the petals. The corolla has 5 petals. The vexillum is the large petal on the posterior-most side. The keel which is covered by the wings of the vexillum is made of the innermost fused petals. The large petals followed by small petals kind of overlapping is seen in the descending order in terms of size. Thus, this arrangement is known as descending imbricate.

7. Flowers are Zygomorphic in (NEET 2011 QP)
a) mustard
b) gulmohur
c) tomato
d) Datura

ANSWER

Flowers of gulmohur have bilateral symmetry. So, they are called zygomorphic. Datura, mustard and tomato have actinomorphic flowers.

8. Epipetalous and syngenesious stamens occur in (NEET 1991 QP)
a) solanaceae
b) Brassicaceae
c) Brassicaceae
d) Asteraceae

ANSWER

Syngenesious condition is found in asteraceae. It is the condition when stamens are united by their anthers (filaments free). Epipetalous condition is also seen here.

9. Which of the following flowers exhibit radial symmetry?
a. Pisum
b. Brassica
c. Cassia
d. Trifolium

ANSWER

Explanation. Out of the given options, Brassica is the only flower which exhibits distinct radial symmetry. All the petals of the flower are identical in size and shape. The other type of symmetry is the bilateral symmetry which shows symmetry in only one plane. The radial symmetry in plants is a pollination strategy.

10) Family Fabaceae differs from Solanaceae and Liliaceae. With respect to the stamens, pick out the characteristics specific to family Fabaceae but not found in Solanaceae or Liliaceae. NEET 2023 QP

a) Polyadelphous and epipetalous stamens
b) Monadelphous and Monothecous anthers
c) Epiphyllous and Dithecous anthers
d) Diadelphous and Dithecous anthers

ANSWER

In the family Fabaceae, the stamens are typically arranged in two groups (diadelphous) and each stamen has two distinct chambers (dithecous) in the anther. This arrangement is not present in Solanaceae or Liliaceae.

11. NEET 2023 QP
Axile placentation is observed in

a) China rose, Beans and Lupin
b) Tomato, Dianthus and Pea
c) China rose, Petunia and Lemon
d) Mustard, Cucumber and Primrose

ANSWER

Axile placentation refers to the arrangement of the placenta at the central axis of the ovary. This pattern can be observed in plants like China rose, Petunia, Lemon, and tomato.

Read more

ROTATIONAL MOTION

We have learnt about various aspects of motion along a straight line: kinematics (where we learned about displacement, velocity, and acceleration), projectile motion (a special case of two-dimensional kinematics), force, and Newton’s laws of motion. Newton’s first law tells us that objects move along a straight line at constant speed unless a net external force acts on them. Therefore, if an object moves along a circular path, it must be experiencing an external force. In this chapter, we explore both circular motion and rotational motion.

Angle of Rotation and Angular Velocity

What exactly do we mean by circular motion or rotation? Rotational motion is the circular motion of an object about an axis of rotation. We will discuss specifically circular motion and spin. Circular motion is when an object moves in a circular path.

Examples of circular motion include a race car speeding around a circular curve, a toy attached to a string swinging in a circle around your head. Spin is rotation about an axis that goes through the center of mass of the object, such as Earth rotating on its axis, a wheel turning on its axle, the spin of a tornado on its path of destruction, or a figure skater spinning during a performance at the Olympics. Sometimes, objects will be spinning while in circular motion, like the Earth spinning on its axis while revolving around the Sun, but we will focus on these two motions separately.

When solving problems involving rotational motion, we use variables that are similar to linear variables (distance, velocity, acceleration, and force) but take into account the curvature or rotation of the motion. Here, we define the angle of rotation, which is the angular equivalence of distance; and angular velocity, which is the angular equivalence of linear velocity.

When objects rotate about some axis—for example, when the CD in the below figure rotates about its center—each point in the object follows a circular path.

All points on a CD travel in circular paths. The pits (dots) along a line from the center to the edge all move through the same angle Δθ in time Δt

The arc length, , is the distance travelled along a circular path. The radius of curvature, r, is the radius of the circular path. Both are shown in the below figure

The radius (r) of a circle is rotated through an angle Δθ . The arc length, ΔS, is the distance covered along the circumference.

Consider a line from the center of the CD to its edge. In a given time, each pit (used to record information) on this line moves through the same angle. The angle of rotation is the amount of rotation and is the angular analog of distance. The angle of rotation Δθ is the arc length divided by the radius of curvature

Δθ = ΔS / r

The angle of rotation is often measured by using a unit called the radian. (Radians are actually dimensionless, because a radian is defined as the ratio of two distances, radius and arc length.) A revolution is one complete rotation, where every point on the circle returns to its original position. One revolution covers 2π radians (or 360 degrees), and therefore has an angle of rotation of 2π radians, and an arc length that is the same as the circumference of the circle. We can convert between radians, revolutions, and degrees using the relationship

1 revolution = 2π rad = 360°. See the below for the conversion of degrees to radians for some common angles

2 π rad = 3600

1 rad = 3600 / 2 π

1 rad ≈ 57.3 0

Degree Measures Radian Measure
300 π/ 6
600 π/ 3
900 π/ 2
1200 2π/ 3
1350 3π/ 4
1800 π

Angular Velocity

How fast is an object rotating? We can answer this question by using the concept of angular velocity. Consider first the angular speed ω is the rate at which the angle of rotation changes. In equation form, the angular speed is

ω = Δθ / Δ t

which means that an angular rotation Δθ occurs in a time, Δt. If an object rotates through a greater angle of rotation in a given time, it has a greater angular speed. The units for angular speed are radians per second (rad/s).

Now let’s consider the direction of the angular speed, which means we now must call it the angular velocity. The direction of the angular velocity is along the axis of rotation. For an object rotating clockwise, the angular velocity points away from you along the axis of rotation. For an object rotating counterclockwise, the angular velocity points toward you along the axis of rotation.

Angular velocity (ω) is the angular version of linear velocity v. Tangential velocity is the instantaneous linear velocity of an object in rotational motion. To get the precise relationship between angular velocity and tangential velocity, consider again a pit on the rotating CD. This pit moves through an arc length  Δs  in a short time  Δt so its tangential speed is

V = Δs / Δt

From the definition of the angle of rotation, we know that

Δϑ = Δs / r,

Hence , Δs = r Δϑ

Substituting this into the expression for v gives

V = r Δϑ / Δt

V = r ω

The equation V = r ω says that the tangential speed v is proportional to the distance r from the center of rotation. Consequently, tangential speed is greater for a point on the outer edge of the CD (with larger r) than for a point closer to the center of the CD (with smaller r). This makes sense because a point farther out from the center has to cover a longer arc length in the same amount of time as a point closer to the center. Note that both points will still have the same angular speed, regardless of their distance from the center of rotation

Points 1 and 2 rotate through the same angle (Δθ), but point 2 moves through a greater arc length (Δs2) because it is farther from the center of rotation.

Now, consider another example: the tire of a moving car . The faster the tire spins, the faster the car moves—large  means large v because . Similarly, a larger-radius tire rotating at the same angular velocity, , will produce a greater linear (tangential) velocity, v, for the car. This is because a larger radius means a longer arc length must contact the road, so the car must move farther in the same amount of time

A car moving at a velocity, v, to the right has a tire rotating with angular velocity ω . The speed of the tread of the tire relative to the axle is v, the same as if the car were jacked up and the wheels spinning without touching the road. Directly below the axle, where the tire touches the road, the tire tread moves backward with respect to the axle with tangential velocity v=rω, where r is the tire radius.

Because the road is stationary with respect to this point of the tire, the car must move forward at the linear velocity v. A larger angular velocity for the tire means a greater linear velocity for the car.

However, there are cases where linear velocity and tangential velocity are not equivalent, such as a car spinning its tires on ice. In this case, the linear velocity will be less than the tangential velocity. Due to the lack of friction under the tires of a car on ice, the arc length through which the tire treads move is greater than the linear distance through which the car moves. It’s similar to running on a treadmill or pedaling a stationary bike; you are literally going nowhere fast

Uniform Circular Motion

In the previous section, we defined circular motion. The simplest case of circular motion is uniform circular motion, where an object travels a circular path at a constant speed. Note that, unlike speed, the linear velocity of an object in circular motion is constantly changing because it is always changing direction. We know from kinematics that acceleration is a change in velocity, either in magnitude or in direction or both. Therefore, an object undergoing uniform circular motion is always accelerating, even though the magnitude of its velocity is constant.

You experience this acceleration yourself every time you ride in a car while it turns a corner. If you hold the steering wheel steady during the turn and move at a constant speed, you are executing uniform circular motion. What you notice is a feeling of sliding (or being flung, depending on the speed) away from the center of the turn. This isn’t an actual force that is acting on you—it only happens because your body wants to continue moving in a straight line (as per Newton’s first law) whereas the car is turning off this straight-line path. Inside the car it appears as if you are forced away from the center of the turn. This fictitious force is known as the centrifugal force. The sharper the curve and the greater your speed, the more noticeable this effect becomes.

Below figure shows an object moving in a circular path at constant speed. The direction of the instantaneous tangential velocity is shown at two points along the path. Acceleration is in the direction of the change in velocity; in this case it points roughly toward the center of rotation. (The center of rotation is at the center of the circular path). If we imagine  becoming smaller and smaller, then the acceleration would point exactly toward the center of rotation, but this case is hard to draw. We call the acceleration of an object moving in uniform circular motion the centripetal acceleration ac because centripetal means center seeking.

The directions of the velocity of an object at two different points are shown, and the change in velocity  is seen to point approximately toward the center of curvature (see small inset). For an extremely small value of Δs, Δv points exactly toward the center of the circle (but this is hard to draw). Because ac = Δv/Δt, the acceleration is also toward the center, so ac is called centripetal acceleration.

Now that we know that the direction of centripetal acceleration is toward the center of rotation, let’s discuss the magnitude of centripetal acceleration. For an object traveling at speed v in a circular path with radius r, the magnitude of centripetal acceleration is

ac  = v2 / r

Centripetal acceleration is greater at high speeds and in sharp curves (smaller radius), as you may have noticed when driving a car, because the car actually pushes you toward the center of the turn. But it is a bit surprising that ac is proportional to the speed squared. This means, for example, that the acceleration is four times greater when you take a curve at 100 km/h than at 50 km/h

We can also express ac in terms of the magnitude of angular velocity. Substituting v = rω into the equation above, we get

ac = (rω)2 / r = rω2

Centripetal Force

Because an object in uniform circular motion undergoes constant acceleration (by changing direction), we know from Newton’s second law of motion that there must be a constant net external force acting on the object.

Any force or combination of forces can cause a centripetal acceleration. Just a few examples are the tension in the rope on a tether ball, the force of Earth’s gravity on the Moon, the friction between a road and the tires of a car as it goes around a curve, or the normal force of a roller coaster track on the cart during a loop-the-loop.

Any net force causing uniform circular motion is called a centripetal force. The direction of a centripetal force is toward the center of rotation, the same as for centripetal acceleration. According to Newton’s second law of motion, a net force causes the acceleration of mass according to Fnet = ma. For uniform circular motion, the acceleration is centripetal acceleration: a = ac.

Therefore, the magnitude of centripetal force, Fc, is Fc = mac

By using the two different forms of the equation for the magnitude of centripetal acceleration, ac=v2/r and ac = rω2, we get two expressions involving the magnitude of the centripetal force Fc. The first expression is in terms of tangential speed, the second is in terms of angular speed:
Fc = m v2/r and Fc = mω2.

Both forms of the equation depend on mass, velocity, and the radius of the circular path. You may use whichever expression for centripetal force is more convenient. Newton’s second law also states that the object will accelerate in the same direction as the net force. By definition, the centripetal force is directed towards the center of rotation, so the object will also accelerate towards the center. A straight line drawn from the circular path to the center of the circle will always be perpendicular to the tangential velocity. Note that, if you solve the first expression for r, you get

r = mv2 / Fc

From this expression, we see that, for a given mass and velocity, a large centripetal force causes a small radius of curvature—that is, a tight curve

In this figure, the frictional force f serves as the centripetal force Fc. Centripetal force is perpendicular to tangential velocity and causes uniform circular motion. The larger the centripetal force Fc, the smaller is the radius of curvature r and the sharper is the curve. The lower curve has the same velocity v, but a larger centripetal force Fc produces a smaller radius r‘.

Rotational Motion

In the section on uniform circular motion, we discussed motion in a circle at constant speed and, therefore, constant angular velocity. However, there are times when angular velocity is not constant—rotational motion can speed up, slow down, or reverse directions. Angular velocity is not constant when a spinning skater pulls in her arms, when a child pushes a merry-go-round to make it rotate, or when a CD slows to a halt when switched off. In all these cases, angular acceleration occurs because the angular velocity ω changes. The faster the change occurs, the greater is the angular acceleration. Angular acceleration α is the rate of change of angular velocity. In equation form, angular acceleration is

α = Δω / Δt

where Δω is the change in angular velocity and Δt is the change in time. The units of angular acceleration are (rad/s)/s, or rad/s2. If ω increases, then α is positive. If ω decreases, then α is negative. Keep in mind that, by convention, counterclockwise is the positive direction and clockwise is the negative direction. For example, a skater rotates counterclockwise as seen from above, so her angular velocity is positive. Acceleration would be negative, for example, when an object that is rotating counterclockwise slows down. It would be positive when an object that is rotating counterclockwise speeds up.

The relationship between the magnitudes of tangential acceleration, a, and angular acceleration,
α, isa = rαorα = a/r.

These equations mean that the magnitudes of tangential acceleration and angular acceleration are directly proportional to each other. The greater the angular acceleration, the larger the change in tangential acceleration, and vice versa.

Rotational  Linear Relationship

θ x θ = x/r
ω v ω = v/r
α a α = a/r

We can now begin to see how rotational quantities like θ, ω and α are related to each other. For example, if a motorcycle wheel that starts at rest has a large angular acceleration for a fairly long time, it ends up spinning rapidly and rotates through many revolutions. Putting this in terms of the variables, if the wheel’s angular acceleration α is large for a long period of time t, then the final angular velocity ω and angle of rotation θ are large. In the case of linear motion, if an object starts at rest and undergoes a large linear acceleration, then it has a large final velocity and will have traveled a large distance.

The kinematics of rotational motion describes the relationships between the angle of rotation, angular velocity, angular acceleration, and time. It only describes motion—it does not include any forces or masses that may affect rotation (these are part of dynamics). Recall the kinematics equation for linear motion:    v = v0 + at (constant a)

As in linear kinematics, we assume a is constant, which means that angular acceleration α is also a constant, because
a = r α

The equation for the kinematics relationship between ω, α, and t is
ω = ω0 + αt (constant α),

where ω0 is the initial angular velocity. Notice that the equation is identical to the linear version, except with angular analogs of the linear variables. In fact, all of the linear kinematics equations have rotational analogs, which are given in Table 6.3. These equations can be used to solve rotational or linear kinematics problem in which a and αare constant.

Torque

If you have ever spun a bike wheel or pushed a merry-go-round, you know that force is needed to change angular velocity. The farther the force is applied from the pivot point (or fulcrum), the greater the angular acceleration. For example, a door opens slowly if you push too close to its hinge, but opens easily if you push far from the hinges. Furthermore, we know that the more massive the door is, the more slowly it opens; this is because angular acceleration is inversely proportional to mass. These relationships are very similar to the relationships between force, mass, and acceleration from Newton’s second law of motion. Since we have already covered the angular versions of distance, velocity and time, you may wonder what the angular version of force is, and how it relates to linear force.

The angular version of force is torque , which is the turning effectiveness of a force. See the below figure. The equation for the magnitude of torque is

where r is the magnitude of the lever arm, F is the magnitude of the linear force, and θ is the angle between the lever arm and the force. The lever arm is the vector from the point of rotation (pivot point or fulcrum) to the location where force is applied. Since the magnitude of the lever arm is a distance, its units are in meters, and torque has units of N⋅m. Torque is a vector quantity and has the same direction as the angular acceleration that it produces.

The harder the man pushes the merry-go- round in the above figure , the faster it accelerates. Furthermore, the more massive the merry-go-round is, the slower it accelerates for the same torque. If the man wants to maximize the effect of his force on the merry-go-round, he should push as far from the center as possible to get the largest lever arm and, therefore, the greatest torque and angular acceleration. Torque is also maximized when the force is applied perpendicular to the lever arm.

Just as linear forces can balance to produce zero net force and no linear acceleration, the same is true of rotational motion. When two torques of equal magnitude act in opposing directions, there is no net torque and no angular acceleration, as you can see in the following video. If zero net torque acts on a system spinning at a constant angular velocity, the system will continue to spin at the same angular velocity

Take Quiz

1) What is the angle of rotation (in degrees) between two hands of a clock, if the radius of the clock is 0.70 m and the arc length separating the two hands is 1.0 m ?

a. 400
b. 800
c. 810
d. 1630

ANSWER

c. 810

2) What is the centripetal force exerted on a 1,600 kg car that rounds a 100 m radius curve at 12 m/s?

a. 192 N
b. 1, 111 N
c. 2, 300 N
d. 13, 333 N

ANSWER

d. 13, 333 N

3) For a given object, what happens to the arc length as the angle of rotation increases?

a. The arc length is directly proportional to the angle of rotation, so it increases with the angle of rotation.
b. The arc length is inversely proportional to the angle of rotation, so it decreases with the angle of rotation.
c. The arc length is directly proportional to the angle of rotation, so it decreases with the angle of rotation.
d. The arc length is inversely proportional to the angle of rotation, so it increases with the angle of rotation

ANSWER

a. The arc length is directly proportional to the angle of rotation, so it increases with the angle of rotation.

4) Which of these quantities is constant in uniform circular motion?

a. Speed
b. Velocity
c. Acceleration
d. Displacement

ANSWER

a. Speed

5) Which of these quantities impact centripetal force?

a. Mass and speed only
b. Mass and radius only
c. Speed and radius only
d. Mass, speed, and radius all impact centripetal force

ANSWER

d. Mass, speed, and radius all impact centripetal force

6) An increase in the magnitude of which of these quantities causes a reduction in centripetal force?

a. Mass
b. Radius of curvature
c. Speed

ANSWER

b. Radius of curvature

7) What happens to centripetal acceleration as the radius of curvature decreases and the speed is constant, and why?

a. It increases, because the centripetal acceleration is inversely proportional to the radius of the curvature.
b. It increases, because the centripetal acceleration is directly proportional to the radius of curvature.
c. It decreases, because the centripetal acceleration is inversely proportional to the radius of the curvature.
d. It decreases, because the centripetal acceleration is directly proportional to the radius of the curvature.

ANSWER

a. It increases, because the centripetal acceleration is inversely proportional to the radius of the curvature.

Read more

LAWS OF MOTION

Motion itself can be beautiful, such as a dolphin jumping out of the water, the flight of a bird, or the orbit of a satellite. The study of motion is called kinematics, but kinematics describes only the way objects move—their velocity and their acceleration. Dynamics considers the forces that affect the motion of moving objects and systems.

Newton’s laws of motion are the foundation of dynamics. These laws describe the way objects speed up, slow down, stay in motion, and interact with other objects. They are also universal laws: they apply everywhere on Earth as well as in space. A force pushes or pulls an object. The object being moved by a force could be an inanimate object, a table, or an animate object, a person. The pushing or pulling may be done by a person, or even the gravitational pull of Earth. Forces have different magnitudes and directions; this means that some forces are stronger than others and can act in different directions.

For example, a cannon exerts a strong force on the cannonball that is launched into the air. In contrast, a mosquito landing on your arm exerts only a small force on your arm. When multiple forces act on an object, the forces combine. Adding together all of the forces acting on an object gives the total force, or net force.

An external force is a force that acts on an object within the system from outside the system. This type of force is different than an internal force, which acts between two objects that are both within the system. The net external force combines these two definitions; it is the total combined external force.

We discuss further details about net force, external force, and net external force in the coming sections. In mathematical terms, two forces acting in opposite directions have opposite signs (positive or negative). By convention, the negative sign is assigned to any movement to the left or downward. If two forces pushing in opposite directions are added together, the larger force will be somewhat cancelled out by the smaller force pushing in the opposite direction. It is important to be consistent with your chosen coordinate system within a problem; for example, if negative values are assigned to the downward direction for velocity, then distance, force, and acceleration should also be designated as being negative in the downward direction.

Newton’s First Law and Friction :

 Newton’s first law of motion states the following:

1. A body at rest tends to remain at rest.

2. A body in motion tends to remain in motion at a constant velocity unless acted on by a net external force. (Recall that constant velocity means that the body moves in a straight line and at a constant speed.)

At first glance, this law may seem to contradict your everyday experience. You have probably noticed that a moving object will usually slow down and stop unless some effort is made to keep it moving. The key to understanding why, for example, a sliding box slows down (seemingly on its own) is to first understand that a net external force acts on the box to make the box slow down. Without this net external force, the box would continue to slide at a constant velocity (as stated in Newton’s first law of motion). What force acts on the box to slow it down? This force is called friction.

Friction is an external force that acts opposite to the direction of motion (see Figure ). Think of friction as a resistance to motion that slows things down.

Consider an air hockey table. When the air is turned off, the puck slides only a short distance before friction slows it to a stop. However, when the air is turned on, it lifts the puck slightly, so the puck experiences very little friction as it moves over the surface. With friction almost eliminated, the puck glides along with very little change in speed. On a frictionless surface, the puck would experience no net external force (ignoring air resistance, which is also a form of friction). Additionally, if we know enough about friction, we can accurately predict how quickly objects will slow down.

Now let’s think about an example. A man pushes a box across a floor at constant velocity by applying a force of +50 N. (The positive sign indicates that, by convention, the direction of motion is to the right.) What is the force of friction that opposes the motion? The force of friction must be −50 N. Why? According to Newton’s first law of motion, any object moving at constant velocity has no net external force acting upon it, which means that the sum of the forces acting on the object F net must be zero. The mathematical way to say that no net external force acts on an object is F net  = 0  or ∑ F = 0 So if the man applies +50 N of force, then the force of friction must be −50 N for the two forces to add up to zero (that is, for the two forces to cancel each other). Whenever you encounter the phrase “at constant velocity” , Newton’s first law tells you that the net external force is zero

The force of friction depends on two factors: the coefficient of friction and the normal force. For any two surfaces that are in contact with one another, the coefficient of friction is a constant that depends on the nature of the surfaces. The normal force is the force exerted by a surface that pushes on an object in response to gravity pulling the object down. In equation form, the force of friction is

f = µ N

Where µ is the coefficient of friction and N is the normal force.

A net external force acts from outside on the object of interest. A more precise definition is that it acts on the system of interest. A system is one or more objects that you choose to study. It is important to define the system at the beginning of a problem to figure out which forces are external and need to be considered, and which are internal and can be ignored.

For example, in the below  Figure , two children push a third child in a wagon at a constant velocity. The system of interest is the wagon plus the small child, as shown in part (b) of the figure. The two children behind the wagon exert external forces on this system (F1, F2). Friction f acting at the axles of the wheels and at the surface where the wheels touch the ground two other external forces acting on the system. Two more external forces act on the system: the weight W of the system pulling down and the normal force N of the ground pushing up. Notice that the wagon is not accelerating vertically, so Newton’s first law tells us that the normal force balances the weight. Because the wagon is moving forward at a constant velocity, the force of friction must have the same strength as the sum of the forces applied by the two children

Figure The wagon and rider form a system that is acted on by external forces. (b) The two children pushing the wagon and child provide two external forces. Friction acting at the wheel axles and on the surface of the tires where they touch the ground provide an external force that act against the direction of motion. The weight W and the normal force N from the ground are two more external forces acting on the system. All external forces are represented in the figure by arrows. All of the external forces acting on the system add together, but because the wagon moves at a constant velocity, all of the forces must add up to zero

Mass and Inertia

Inertia is the tendency for an object at rest to remain at rest, or for a moving object to remain in motion in a straight line with constant speed. This key property of objects was first described by Galileo. Later, Newton incorporated the concept of inertia into his first law, which is often referred to as the law of inertia. As we know from experience, some objects have more inertia than others. For example, changing the motion of a large truck is more difficult than changing the motion of a toy truck. In fact, the inertia of an object is proportional to the mass of the object. Mass is a measure of the amount of matter (or stuff) in an object. The quantity or amount of matter in an object is determined by the number and types of atoms the object contains. Unlike weight (which changes if the gravitational force changes), mass does not depend on gravity. The mass of an object is the same on Earth, in orbit, or on the surface of the moon. In practice, it is very difficult to count and identify all of the atoms and molecules in an object, so mass is usually not determined this way. Instead, the mass of an object is determined by comparing it with the standard kilogram. Mass is therefore expressed in kilograms.

Newton’s Second Law of Motion :

Newton’s first law considered bodies at rest or bodies in motion at a constant velocity. The other state of motion to consider is when an object is moving with a changing velocity, which means a change in the speed and/or the direction of motion. This type of motion is addressed by Newton’s second law of motion, which states how force causes changes in motion. Newton’s second law of motion is used to calculate what happens in situations involving forces and motion, and it shows the mathematical relationship between force, mass, and acceleration. Mathematically, the second law is most often written as

Fnet = m a or ∑ F = m a

where Fnet  (or ∑F) is the net external force, m is the mass of the system, and a is the acceleration. Note that Fnet and ∑F are the same because the net external force is the sum of all the external forces acting on the system.

what do we mean by a change in motion? A change in motion is simply a change in velocity: the speed of an object can become slower or faster, the direction in which the object is moving can change, or both of these variables may change. A change in velocity means, by definition, that an acceleration has occurred.

Newton’s first law says that only a nonzero net external force can cause a change in motion, so a net external force must cause an acceleration. Note that acceleration can refer to slowing down or to speeding up. Acceleration can also refer to a change in the direction of motion with no change in speed, because acceleration is the change in velocity divided by the time it takes for that change to occur, and velocity is defined by speed and direction. From the equation we see that force is directly proportional to both mass and acceleration, which makes sense. To accelerate two objects from rest to the same velocity, you would expect more force to be required to accelerate the more massive object. Likewise, for two objects of the same mass, applying a greater force to one would accelerate it to a greater velocity. Now, let’s rearrange Newton’s second law to solve for acceleration. We get

 a = Fnet / m or a = ∑F / m

In this form, we can see that acceleration is directly proportional to force, which we write as

a α Fnet

where the symbol means proportional to. This proportionality mathematically states what we just said in words: acceleration is directly proportional to the net external force. When two variables are directly proportional to each other, then if one variable doubles, the other variable must double. Likewise, if one variable is reduced by half, the other variable must also be reduced by half. In general, when one variable is multiplied by a number, the other variable is also multiplied by the same number. It seems reasonable that the acceleration of a system should be directly proportional to and in the same direction as the net external force acting on the system. An object experiences greater acceleration when acted on by a greater force

It is also clear from the equation  a = Fnet / m , that acceleration is inversely proportional to mass, which we write as

a α 1/m

Inversely proportional means that if one variable is multiplied by a number, the other variable must be divided by the same number. Now, it also seems reasonable that acceleration should be inversely proportional to the mass of the system. In other words, the larger the mass (the inertia), the smaller the acceleration produced by a given force. This relationship is illustrated in Figure 4.5, which shows that a given net external force applied to a basketball produces a much greater acceleration than when applied to a car.

Applying Newton’s Second Law :

Before putting Newton’s second law into action, it is important to consider units. The equation

Fnet = m a is used to define the units of force in terms of the three basic units of mass, length, and time (recall that acceleration has units of length divided by time squared). The SI unit of force is called the newton (abbreviated N) and is the force needed to accelerate a 1-kg system at the rate of 1 m/s 2. That is, because Fnet = m a  we have

1 N = 1 Kg X 1 m/s2  = 1 Kg m s -2

One of the most important applications of Newton’s second law is to calculate weight (also known as the gravitational force), which is usually represented mathematically as W. When people talk about gravity, they don’t always realize that it is an acceleration. When an object is dropped, it accelerates toward the center of Earth. Newton’s second law states that the net external force acting on an object is responsible for the acceleration of the object. If air resistance is negligible, the net external force on a falling object is only the gravitational force (i.e., the weight of the object).

 Weight can be represented by a vector because it has a direction. Down is defined as the direction in which gravity pulls, so weight is normally considered a downward force. By using Newton’s second law, we can figure out the equation for weight.

Consider an object with mass m falling toward Earth. It experiences only the force of gravity (i.e., the gravitational force or weight), which is represented by W. Newton’s second law states that  Because the only force acting on the object is the gravitational force, we have

Fnet = W 

We know that the acceleration of an object due to gravity is g, so we have

a = g

Substituting these two expressions into Newton’s second law gives

W = mg

This is the equation for weight—the gravitational force on a mass m. On Earth, g = 9.8 m/s2  so the weight (disregarding for now the direction of the weight) of a 1.0-kg object on Earth is

W = mg = 1 X 9.8 m/s2 = 9.8 N

When the net external force on an object is its weight, we say that it is in freefall. In this case, the only force acting on the object is the force of gravity.

 On the surface of Earth, when objects fall downward toward Earth, they are never truly in freefall because there is always some upward force due to air resistance that acts on the object (and there is also the buoyancy force of air, which is similar to the buoyancy force in water that keeps boats afloat).

Gravity varies slightly over the surface of Earth, so the weight of an object depends very slightly on its location on Earth. Weight varies dramatically away from Earth’s surface. On the moon, for example, the acceleration due to gravity is only 1.67 m/s2 . Because weight depends on the force of gravity, a 1.0-kg mass weighs 9.8 N on Earth and only about 1.7 N on the moon.

It is important to remember that weight and mass are very different, although they are closely related. Mass is the quantity of matter (how much stuff) in an object and does not vary, but weight is the gravitational force on an object and is proportional to the force of gravity.

It is easy to confuse the two, because our experience is confined to Earth, and the weight of an object is essentially the same no matter where you are on Earth. Adding to the confusion, the terms mass and weight are often used interchangeably in everyday language; for example, our medical records often show our weight in kilograms, but never in the correct unit of newtons.

Momentum, Impulse, and the Impulse-Momentum Theorem :

Linear momentum is the product of a system’s mass and its velocity. In equation form, linear momentum p is

P = m V

You can see from the equation that momentum is directly proportional to the object’s mass (m) and velocity (v). Therefore, the greater an object’s mass or the greater its velocity, the greater its momentum.

A large, fast-moving object has greater momentum than a smaller, slower object. Momentum is a vector and has the same direction as velocity v. Since mass is a scalar, when velocity is in a negative direction (i.e., opposite the direction of motion), the momentum will also be in a negative direction; and when velocity is in a positive direction, momentum will likewise be in a positive direction. The SI unit for momentum is kg m/s.

Momentum is so important for understanding motion that it was called the quantity of motion by physicists such as Newton. Force influences momentum, and we can rearrange Newton’s second law of motion to show the relationship between force and momentum.

Newton actually stated his second law of motion in terms of momentum: The net external force equals the change in momentum of a system divided by the time over which it changes. The change in momentum is the difference between the final and initial values of momentum. In equation form, this law is

Fnet = ΔP / Δt

where Fnet is the net external force, ΔP is the change in momentum, and Δt  is the change in time. We can solve for ΔP  by rearranging the equation

ΔP = Fnet Δt

Fnet Δt is known as impulse and this equation is known as the impulse-momentum theorem.

From the equation, we see that the impulse equals the average net external force multiplied by the time this force acts. It is equal to the change in momentum. The effect of a force on an object depends on how long it acts, as well as the strength of the force. Impulse is a useful concept because it quantifies the effect of a force. A very large force acting for a short time can have a great effect on the momentum of an object, such as the force of a racket hitting a tennis ball. A small force could cause the same change in momentum, but it would have to act for a much longer time

Newton’s Second Law in Terms of Momentum

When Newton’s second law is expressed in terms of momentum, it can be used for solving problems where mass varies, since ΔP = Δ(mv)  . In the more traditional form of the law that you are used to working with, mass is assumed to be constant. In fact, this traditional form is a special case of the law, where mass is constant.Fnet = ma  is actually derived from the equation:

Fnet = ΔP / Δt

The change in momentum ΔP  is given by

ΔP = Δ (mv)

If the mass of the system is constant, then

Δ (mv) = m Δv

By substituting m Δv for ΔP , Newton’s second law of motion becomes

Fnet = ΔP / Δt

      = m Δv / Δt for a constant mass

Since Δv / Δt = a , we get

Fnet = ma , when the mass of the system is constant

Conservation of Momentum

It is important we realize that momentum is conserved during collisions, explosions, and other events involving objects in motion. To say that a quantity is conserved means that it is constant throughout the event. In the case of conservation of momentum, the total momentum in the system remains the same before and after the collision.

You may have noticed that momentum was not conserved , where forces acting on the objects produced large changes in momentum. Why is this? The systems of interest considered in those problems were not inclusive enough. If the systems were expanded to include more objects, then momentum would in fact be conserved in those sample problems. It is always possible to find a larger system where momentum is conserved, even though momentum changes for individual objects within the system. For example, if a football player runs into the goalpost in the end zone, a force will cause him to bounce backward. His momentum is obviously greatly changed, and considering only the football player, we would find that momentum is not conserved. However, the system can be expanded to contain the entire Earth. Surprisingly, Earth also recoils—conserving momentum—because of the force applied to it through the goalpost. The effect on Earth is not noticeable because it is so much more massive than the player, but the effect is real. Next, consider what happens if the masses of two colliding objects are more similar than the masses of a football player and Earth—in the example shown in Figure 8.4 of one car bumping into another. Both cars are coasting in the same direction when the lead car, labeled m2 , is bumped by the trailing car, labeled m1 . The only unbalanced force on each car is the force of the collision, assuming that the effects due to friction are negligible. Car m1 slows down as a result of the collision, losing some momentum, while car m2 speeds up and gains some momentum. If we choose the system to include both cars and assume that friction is negligible, then the momentum of the two-car system should remain constant. Now we will prove that the total momentum of the two-car system does in fact remain constant, and is therefore conserved

Newton’s Third Law of Motion :

Newton’s third law of motion states that whenever a first object exerts a force on a second object, the first object experiences a force equal in magnitude but opposite in direction to the force that it exerts.

Newton’s third law of motion tells us that forces always occur in pairs, and one object cannot exert a force on another without experiencing the same strength force in return. We sometimes refer to these force pairs as action-reaction pairs, where the force exerted is the action, and the force experienced in return is the reaction (although which is which depends on your point of view).

Newton’s third law is useful for figuring out which forces are external to a system. Recall that identifying external forces is important when setting up a problem, because the external forces must be added together to find the net force.

We can see Newton’s third law at work by looking at how people move about. Consider a swimmer pushing off from the side of a pool. He pushes against the pool wall with his feet and accelerates in the direction opposite to the push. The wall has thus exerted on the swimmer a force of equal magnitude but in the direction opposite that of his push. You might think that two forces of equal magnitude but that act in opposite directions would cancel, but they do not because they act on different systems.

Similarly, a car accelerates because the ground pushes forward on the car’s wheels in reaction to the car’s wheels pushing backward on the ground. You can see evidence of the wheels pushing backward when tires spin on a gravel road and throw rocks backward.

Birds fly by exerting force on air in the direction opposite that in which they wish to fly. For example, the wings of a bird force air downward and backward in order to get lift and move forward

Applying Newton’s Third Law :

The gravitational force (or weight) acts on objects at all times and everywhere on Earth. We know from Newton’s second law that a net force produces an acceleration; so, why is everything not in a constant state of freefall toward the center of Earth? The answer is the normal force. The normal force is the force that a surface applies to an object to support the weight of that object; it acts perpendicular to the surface upon which the object rests. If an object on a flat surface is not accelerating, the net external force is zero, and the normal force has the same magnitude as the weight of the system but acts in the opposite direction. In equation form, we write that

N = mg

Note that this equation is only true for a horizontal surface.

The word tension comes from the Latin word meaning to stretch. Tension is the force along the length of a flexible connector, such as a string, rope, chain, or cable. Regardless of the type of connector attached to the object of interest, one must remember that the connector can only pull (or exert tension) in the direction parallel to its length. Tension is a pull that acts parallel to the connector, and that acts in opposite directions at the two ends of the connector. This is possible because a flexible connector is simply a long series of action-reaction forces, except at the two ends where outside objects provide one member of the actionreaction forces.

Consider a person holding a mass on a rope, as shown in below Figure

Tension in the rope must equal the weight of the supported mass, as we can prove by using Newton’s second law. If the 5 kg mass in the figure is stationary, then its acceleration is zero, so         Fnet = 0 .  The only external forces acting on the mass are its weight W and the tension T supplied by the rope. Summing the external forces to find the net force, we obtain

Fnet  = T – W = 0

where T and W are the magnitudes of the tension and weight, respectively, and their signs indicate direction, with up being positive. By substituting mg for Fnet and rearranging the equation, the tension equals the weight of the supported mass, we get

T = W = mg

For a 5 Kg mass , the we see that

T = mg = 5 Kg X 9.8 m/s2 = 49 N

Another example of Newton’s third law in action is thrust. Rockets move forward by expelling gas backward at a high velocity. This means that the rocket exerts a large force backward on the gas in the rocket combustion chamber, and the gas, in turn, exerts a large force forward on the rocket in response. This reaction force is called thrust.

Take Quiz

1. An object is at rest. Two forces, X and Y, are acting on it. Force X has a magnitude of x and acts in the downward direction. What is the magnitude and direction of Y?

a. The magnitude is x and points in the upward direction.
b. The magnitude is 2x and points in the upward direction.
c. The magnitude is x and points in the downward direction.
d. The magnitude is 2x and points in the downward direction

ANSWER

a. The magnitude is x and points in the upward direction.

2) Three forces, A, B, and C, are acting on the same object with magnitudes a, b, and c, respectively. Force A acts to the right, force B acts to the left, and force C acts downward. What is a necessary condition for the object to move straight down?

a. The magnitude of force A must be greater than the magnitude of force B, so a > b
b. The magnitude of force A must be equal to the magnitude of force B, so a = b
c. The magnitude of force A must be greater than the magnitude of force C, so A > C.
d. The magnitude of force C must be greater than the magnitude of forces A or B, so A < C > B

ANSWER

d. The magnitude of force C must be greater than the magnitude of forces A or B, so A < C > B

3. Two people push a cart on a horizontal surface by applying forces F1 and F2 in the same direction. Is the magnitude of the net force acting on the cart, Fnet , equal to, greater than, or less than F1 + F2 ? Why?

a. Fnet < F1 + F2 because the net force will not include the frictional force.
b. Fnet = F1 + F2 because the net force will not include the frictional force
c. Fnet  < F1 + F2 because the net force will include the component of frictional force
d. Fnet = F1 + F2 because the net force will include the frictional force

ANSWER

b. Fnet = F1 + F2 because the net force will not include the frictional force

4) True or False: A book placed on a balance scale is balanced by a standard 1-kg iron weight placed on the opposite side of the balance. If these objects are taken to the moon and a similar exercise is performed, the balance is still level because gravity is uniform on the moon’s surface as it is on Earth’s surface.

a. True
b. False

ANSWER

b. False

5) A car weighs 2,000 kg. It moves along a road by applying a force on the road with a parallel component of 560 N. There are two passengers in the car, each weighing 55 kg. If the magnitude of the force of friction experienced by the car is 45 N, what is the acceleration of the car?

a. 0.244 m/s2
b. 0.265 m/s2
c. 4.00 m/s2
d. 4.10 m/s2

ANSWER

a. 0.244 m/s2

6. A person pushes an object of mass 5.0 kg along the floor by applying a force. If the object experiences a friction force of 10 N and accelerates at 18 m/s2 , what is the magnitude of the force exerted by the person?

a. −90 N
b. −80 N
c. 90 N
d. 100 N

ANSWER

d. 100 N

7. A 2,000-kg car is sitting at rest in a parking lot. A bike and rider with a total mass of 60 kg are traveling along a road at 10 km/h. Which system has more inertia? Why?

a. The car has more inertia, as its mass is greater than the mass of the bike.
b. The bike has more inertia, as its mass is greater than the mass of the car.
c. The car has more inertia, as its mass is less than the mass of the bike.
d. The bike has more inertia, as its mass is less than the mass of the car

ANSWER

a. The car has more inertia, as its mass is greater than the mass of the bike.

8. Two people push a 2,000-kg car to get it started. An acceleration of at least 5.0 m/s2 is required to start the car. Assuming both people apply the same magnitude force, how much force will each need to apply if friction between the car and the road is 300 N?

a. 4850 N
b. 5150 N
c. 97000 N
d. 10300 N

ANSWER

b. 5150 N

9. A 55-kg lady stands on a bathroom scale inside an elevator. The scale reads 70 kg. What do you know about the motion of the elevator?

a. The elevator must be accelerating upward.
b. The elevator must be accelerating downward.
c. The elevator must be moving upward with a constant velocity.
d. The elevator must be moving downward with a constant velocity.

ANSWER

a. The elevator must be accelerating upward.

10. Are rockets more efficient in Earth’s atmosphere or in outer space? Why?

a. Rockets are more efficient in Earth’s atmosphere than in outer space because the air in Earth’s atmosphere helps to provide thrust for the rocket, and Earth has more air friction than outer space.
b. Rockets are more efficient in Earth’s atmosphere than in outer space because the air in Earth’s atmosphere helps to provide thrust to the rocket, and Earth has less air friction than the outer space.
c. Rockets are more efficient in outer space than in Earth’s atmosphere because the air in Earth’s atmosphere does not provide thrust but does create more air friction than in outer space.
d. Rockets are more efficient in outer space than in Earth’s atmosphere because the air in Earth’s atmosphere does not provide thrust but does create less air friction than in outer space.

ANSWER

c. Rockets are more efficient in outer space than in Earth’s atmosphere because the air in Earth’s atmosphere does not provide thrust but does create more air friction than in outer space.

Read more

Morphology of flowering plants

Floral Formula

Definition of Floral Formula

A floral formula can defined as the numeric and symbolic expression, which reveals the flower morphological characteristics by employing different symbols, letters, and figures. It is the conventional method accustomed to formulating the structure of the flower. It elucidates the information about the number of whorls and a relative relationship between each other.

Components

The floral formula uses discrete letters, signs and figures to represent the specific feature of the flower.

Letters used in Floral Formula

1.K: This letter denotes the sepals that form an outermost whorl called the calyx.

2. C: This letter represents the group of petals that constitute the second whorl called the corolla.

3. P: It is used to denote the tepals, which indicates the undifferentiated condition of the perianth members (sepals and petals).

4. A: It specifies the male reproduction part or stamens, which includes the filament and anther that all together makes up the third floral whorl known as androecium.

5. G: It denotes the female reproductive part, i.e. carpel, which includes the stigma, style and ovary that colloquially forms the innermost whorl called gynoecium.

6. Br: It represents the bracteate condition of the flower.

A plant having bracts is referred to as bracteate or bracteolate, while one that lacks them is referred to as ebracteate and ebracteolate, without bracts.
bracteate flowers: Flowers with bracts (a reduced leaf at the base of the pedicel) are called bracteate flowers. Bracts are small leaf-like structures found at a flower’s base. China rose, tulip, lily

7. Ebr: It indicates the ebracteate condition, in which a flower lacks bract.

8. Brl: It indicates the presence of bracteoles or represents the bracteolate condition.

9. Epik: It represents the presence of a secondary whorl surrounding the calyx called epicalyx.

10. Ebrl: It is used to indicate the absence of bracteoles or to represent the ebracteolate condition.

Symbols used in Floral Formula

  1. 0: It indicates the absence of a particular member in flower.
  2. : It is generally used when the number of specific flower parts is more than 10.
  3. : It indicates an actinomorphic condition.
  4. %: It denotes a zygomorphic condition.

  • : It represents the bisexuality of flowers.
  • : It represents the unisexual, staminate flower.
  • ♀️: This represents the unisexual, pistillate flower.

Combination of Letters, Symbols and Numbers

  • K5: It shows the aposepalous condition, in which the five sepals are free.

  • K(5): It shows the gamosepalous condition, in which the five sepals are united.

  • C5: It represents the apopetalous condition or the presence of the five free petals.

  • C(5): It represents the gamopetalous condition or the presence of five fused petals.

  • Cx: It indicates corolla cruciform.
  • curve drawn over the letters P and A: It represents the epiphyllous stamens.
  • curve drawn over the letters C and A: It shows the epipetalous stamens.
  • A3: It indicates the presence of three free stamens.
  • A2+2: It indicates the presence of  4 free stamens, two in each whorl.
  • A(9)+1: It represents the presence of diadelphous stamens (10 in number), in which nine stamens are fused in one whorl, and one stamen remains free.

  • A0: Sterile stamen (staminode).
  • G0: Sterile carpel (pistillode).
  • G-: It represents the semi inferior ovary.
  • line over the letter G: It represents the inferior position of an ovary.
  • The line below the letter G: It represents the superior position of an ovary.
  • curve over the letters G and A: It shows the gynostagium condition.

Floral Formula:

Symbol Description / Full form
Br Bracteate condition
Ebr Ebracteate condition
Brl Bracteolate
Epik Epicalyx
Ebrl Ebracteolate
0 (zero) Absence of a particular whorl
Indefinite number of floral parts in a whorl
Actinomorphic condition (radially symmetric) flowers.
% Zygomorphic condition (bilaterally symmetric)
K Calyx (Sepals)
C Corolla (petals)
P Perianth (tepals)
A Androecium (stamens)
G Gynoecium (carpels)
Bisexual flower
Unisexual, staminate flower
♀️ Unisexual, pistillate flower
K5 Five sepals, aposepalous
K(5) Five sepals gamosepalous
C5 Five petals, apopetalous
C(5) Five petals, gamopetalous
Epiphyllous stamens
Epipetalous stamens
A3 Three stamens free
A2+2 Stamens 4, 2 whorls
A10+1 Stamens 10, diadelphous – 9 stamens unite to form one bundle and 1 other bundle
A0 Sterile stamen (staminode)
G0 Sterile carpel (pistillode)
G- Semi inferior ovary
This image has an empty alt attribute; its file name is 3.png Inferior ovary
This image has an empty alt attribute; its file name is 4.png Superior ovary

How to Use Floral Formula?

Firstly, the presence or absence of bract and bracteoles in the flowers of different families is observed and noted by using particular letters.

Then the symmetry of the flower is written according to the position of floral whorls relative to the mother axis. Flowers can be polysymmetric (actinomorphic), dissymmetric, monosymmetric (zygomorphic), asymmetric or spirally arranged. Therefore, the flowers showing different symmetry is indicated by specific symbols in the floral formula.

After that, the sex of the flower is written, i.e. whether a plant is hermaphrodite (bisexual) or unisexual. Bisexual flowers possess the male plus female reproductive parts, while the unisexual flowers either contain stamens or carpels.

Then the number of floral members starting with calyx, corolla, androecium to gynoecium is written. Here each member is denoted by using specific letters like ‘K’ represents calyx, ‘C’ represents corolla, ‘A’ indicates androecium and ‘G’ represents gynoecium. The number of each whorl is written after the following letters starting from K, C, A to G.

The number used after the letter is placed inside bracket ( ), If the particular whorl is united. But, if the floral members remain free, then the bracket is not applied. Adnation between the floral whorls is represented by drawing a curve from the top of the letters.

The position of the ovary can be shown by drawing a horizontal line above, below or in front of the letter ‘G’, which represents the inferior, superior or half inferior position of the ovary.

Examples

The floral formula can be used to describe the flowers of the particular family or the different species of flower. Let us take few examples of the floral formula to study the floral characteristics.

Floral Formula of Fabaceae Family

(%) shows that the symmetry of the flower is monosymmetric or zygomorphic. (⚥) indicates that the flower is perfect or hermaphrodite (includes stamen plus pistil). A letter ‘K’ indicates the outermost whorl, i.e. calyx and (5) shows that the number of sepals is five that are united with each other.

‘C’ indicates the second floral whorl, i.e. corolla or petals and the number after this, i.e. 1+2+(2), represents vexillary aestivation. ‘A’ means the male reproductive part or androecium, and the number after it (9)+1 shows the presence of diadelphous stamens (includes 10 stamens).

In 10 stamens, 9 filaments unite to constitute one bundle, and the remaining one filament makes up another one. The line below the letter ‘G’ indicates the superior position of an ovary, and digit 1 shows the monocarpellary ovary.

Floral Formula of Liliaceae Family

Through this floral formula, we can identify many floral characteristics. This formula indicates that the flower is bracteate, bisexual or hermaphrodite. A letter ‘P’ indicates that the sepals and petals are undifferentiated, i.e. there are three tepals in every two whorls.

Letter ‘A’ denotes androecium, and ‘3+3’ suggests there are three free stamens in every two whorls. A line below a letter ‘G’ specifies the hypogynous condition of an ovary, and the number ‘(3)’ represents the presence of three fused carpels.

Floral Formula of Solanacae Family

The general floral formula of Solanaceae family is as follows:

Here the symbols represent:

Actinomorphic (radial symmetry)
Bisexual
K(5) Calyx – 5 sepals, gamosepalous (united)
C(5) Corolla – 5 petals, gamopetalous
A5 Androecium – 5 stamens, polyandrous (free), epipetalous (attached to petals)
G(2) Gynoecium – bicarpellary, syncarpous (united), superior ovary

  • Calyx (K): Composed of 5 sepals, which are united (gamosepalous).
  • Corolla ©: Consists of 5 petals, which are also united (gamopetalous).
  • Androecium (A): Contains 5 stamens, which are free (polyandrous) and attached to the petals (epipetalous).
  • Gynoecium (G): Bicarpellary (two carpels), syncarpous (united), and has a superior ovary.

The Solanaceae family includes a variety of plants, such as potatoes, tomatoes, bell peppers, eggplants, and tobacco. These plants exhibit diverse characteristics and are widely distributed across tropical, subtropical, and temperate zones.

Fabaceae floral formula

Take Quiz

1. NEET 2021 QP
Match Column – I with Column – II

Select the correct answer from the options given below.

  1. (a)-(iv), (b)-(ii), (c)-(i), (d)-(iii)
  2. (a)-(iii), (b)-(iv), (c)-(ii), (d)-(i)
  3. (a)-(i), (b)-(ii), (c)-(iii), (d)-(iv)
  4. (a)-(ii), (b)-(iii), (c)-(iv), (d)-(i)

ANSWER

b) (a)-(iii), (b)-(iv), (c)-(ii), (d)-(i)

Fabaceae: The flowers in Fabaceae exhibit zygomorphic symmetry and are bisexual. The calyx is composed of five fused sepals, while the corolla is papilionaceous, consisting of one standard petal, two wing petals, and two fused keel petals. The stamens are diadelphous, and the gynoecium is monocarpellary with a superior ovary.

Solanaceae: Solanaceae flowers are actinomorphic and bisexual. The calyx features five fused sepals, and the corolla consists of five fused petals. There are five epipetalous stamens, and the gynoecium is bicarpellary and apocarpous with a superior ovary.
Liliaceae: Liliaceae flowers are actinomorphic and bisexual. The perianth comprises six fused tepals, arranged in two rows. There are six epipetalous stamens, and the gynoecium is tricarpellary and syncarpous with a superior ovary.

Brassicaceae: Flowers in Brassicaceae exhibit actinomorphic symmetry and are bisexual. The calyx consists of four sepals arranged in two rows (polysepalous), while the corolla comprises four petals (polypetalous). There are six stamens, arranged in two rows (tetradynamous), and the gynoecium is bicarpellary and syncarpous with a superior ovary.

So, the correct option is (B): (a)-(iii), (b)-(iv), (c)-(ii), (d)-(i)

2) Floral formula below belongs to which family?

  1. Solanaceae
  2. Liliaceae
  3. Fabaceae
  4. Brassicaceae

ANSWER

c) Fabaceae

(%) shows that the symmetry of the flower is monosymmetric or zygomorphic. (⚥) indicates that the flower is perfect or hermaphrodite (includes stamen plus pistil). A letter ‘K’ indicates the outermost whorl, i.e. calyx and (5) shows that the number of sepals is five that are united with each other.

‘C’ indicates the second floral whorl, i.e. corolla or petals and the number after this, i.e. 1+2+(2), represents vexillary aestivation. ‘A’ means the male reproductive part or androecium, and the number after it (9)+1 shows the presence of diadelphous stamens (includes 10 stamens).

In 10 stamens, 9 filaments unite to constitute one bundle, and the remaining one filament makes up another one. The line below the letter ‘G’ indicates the superior position of an ovary, and digit 1 shows the monocarpellary ovary.

3) Which of the following is false about floral formula – 

  1. Indefinite number of floral parts is shown with the help of infinity symbol ‘∞’
  2. Calyx is denoted with the letter ‘C’
  3. K5 means there are five petals, aposepalous
  4. A2+2 means 4 free stamens, two in each whorl

ANSWER

b) Calyx is denoted with the letter ‘C’

 In the floral formula, calyx (sepals) is denoted as ‘K’ and not ‘C’. it is one of the non-essential organs of the flower, it is the outermost and the first whorl of the flower. In the bud stage of the flower the calyx forms a protective layer around the bud. The ‘C’ in the floral formula is written for corolla.
K5: It shows the aposepalous condition, in which the five sepals are free.
K(5): It shows the gamosepalous condition, in which the five sepals are united.
C5: It represents the apopetalous condition or the presence of the five free petals.
C(5): It represents the gamopetalous condition or the presence of five fused petals.

4) The correct floral formula of soyabean is (NEET 2010 QP)

ANSWER

The plants belonging to the family fabaceae such as soyabean, pea, moong, gram, etc have the floral formula

5) NEET 2016 QP
Tricarpellary, syncarpous gynoecium is found in the flowers of
(a) Liliaceae
(b) Poaceae
(c) Fabaceae
(d) Solanaceae

ANSWER

a) Tricarpellary, syncarpous gynoecium is a unique reproductive structure found in the flowers of Liliaceae family. It is formed by the fusion of three carpels, resulting in a single, compound ovary with three chambers. Each chamber has its own style and stigma, which are designed to facilitate pollination and fertilization.

Read more

Work, Energy, Power and Collision

Work, Power, and the Work–Energy Theorem

In physics, the term work has a very specific definition. Work is application of force, f, to move an object over a distance, d, in the direction that the force is applied. Work, W, is described by the equation

W = fd

Some things that we typically consider to be work are not work in the scientific sense of the term. Let’s consider a few examples. Think about why each of the following statements is true.

  • Homework is not work.
  • Lifting a rock upwards off the ground is work.
  • Carrying a rock in a straight path across the lawn at a constant speed is not work.

The first two examples are fairly simple. Homework is not work because objects are not being moved over a distance. Lifting a rock up off the ground is work because the rock is moving in the direction that force is applied. The last example is less obvious. From the laws of motion , we know that  that force is not required to move an object at constant velocity. Therefore, while some force may be applied to keep the rock up off the ground, no net force is applied to keep the rock moving forward at constant velocity.

Work and energy are closely related. When you do work to move an object, you change the object’s energy. You (or an object) also expend energy to do work. In fact, energy can be defined as the ability to do work. Energy can take a variety of different forms, and one form of energy can transform to another. In this chapter we will be concerned with mechanical energy, which comes in two forms: kinetic energy and potential energy.

  • Kinetic energy is also called energy of motion. A moving object has kinetic energy.
  • Potential energy, sometimes called stored energy, comes in several forms. Gravitational potential energy is the stored energy an object has as a result of its position above Earth’s surface (or another object in space). A roller coaster car at the top of a hill has gravitational potential energy.

Let’s examine how doing work on an object changes the object’s energy. If we apply force to lift a rock off the ground, we increase the rock’s potential energy, PE. If we drop the rock, the force of gravity increases the rock’s kinetic energy as the rock moves downward until it hits the ground.

The force we exert to lift the rock is equal to its weight, w, which is equal to its mass, m, multiplied by acceleration due to gravity, g.

The work we do on the rock equals the force we exert multiplied by the distance, d, that we lift the rock. The work we do on the rock also equals the rock’s gain in gravitational potential energy, PEe.

Kinetic energy depends on the mass of an object and its velocity, v.

When we drop the rock the force of gravity causes the rock to fall, giving the rock kinetic energy. When work done on an object increases only its kinetic energy, then the net work equals the change in the value of the quantity 1/2 mv2. This is a statement of the work–energy theorem, which is expressed mathematically as

The subscripts 2 and 1 indicate the final and initial velocity, respectively. This theorem was proposed and successfully tested by James Joule

The joule (J) is the metric unit of measurement for both work and energy. The measurement of work and energy with the same unit reinforces the idea that work and energy are related and can be converted into one another. 1.0 J = 1.0 N∙m, the units of force multiplied by distance. 1.0 N = 1.0 k∙m/s2, so 1.0 J = 1.0 k∙m2/s2. Analyzing the units of the term (1/2)mv2 will produce the same units for joules.

Calculations Involving Work and Power

In applications that involve work, we are often interested in how fast the work is done. For example, in roller coaster design, the amount of time it takes to lift a roller coaster car to the top of the first hill is an important consideration. Taking a half hour on the ascent will surely irritate riders and decrease ticket sales. Let’s take a look at how to calculate the time it takes to do work.

Recall that a rate can be used to describe a quantity, such as work, over a period of time. Power is the rate at which work is done. In this case, rate means per unit of time. Power is calculated by dividing the work done by the time it took to do the work.

P = W / t

Power can be expressed in units of watts (W). This unit can be used to measure power related to any form of energy or work. You have most likely heard the term used in relation to electrical devices, especially light bulbs. Multiplying power by time gives the amount of energy. Electricity is sold in kilowatt-hours because that equals the amount of electrical energy consumed.

Mechanical Energy and Conservation of Energy

We saw earlier that mechanical energy can be either potential or kinetic. In this section we will see how energy is transformed from one of these forms to the other. We will also see that, in a closed system, the sum of these forms of energy remains constant.

Imagine a roller coaster ride . Quite a bit of potential energy is gained by a roller coaster car and its passengers when they are raised to the top of the first hill. Remember that the potential part of the term means that energy has been stored and can be used at another time. You will see that this stored energy can either be used to do work or can be transformed into kinetic energy. For example, when an object that has gravitational potential energy falls, its energy is converted to kinetic energy. Remember that both work and energy are expressed in joules.

Now, let’s look at the roller coaster in the below figure . Work was done on the roller coaster to get it to the top of the first rise; at this point, the roller coaster has gravitational potential energy. It is moving slowly, so it also has a small amount of kinetic energy. As the car descends the first slope, its PE is converted to KE. At the low point much of the original PE has been transformed to KE, and speed is at a maximum. As the car moves up the next slope, some of the KE is transformed back into PE and the car slows down.

On an actual roller coaster, there are many ups and downs, and each of these is accompanied by transitions between kinetic and potential energy. Assume that no energy is lost to friction. At any point in the ride, the total mechanical energy is the same, and it is equal to the energy the car had at the top of the first rise. This is a result of the law of conservation of energy, which says that, in a closed system, total energy is conserved—that is, it is constant. Using subscripts 1 and 2 to represent initial and final energy, this law is expressed as

Either side equals the total mechanical energy. The phrase in a closed system means we are assuming no energy is lost to the surroundings due to friction and air resistance. If we are making calculations on dense falling objects, this is a good assumption.

Elastic and Inelastic Collisions

When objects collide, they can either stick together or bounce off one another, remaining separate. In this section, we’ll cover these two different types of collisions, first in one dimension and then in two dimensions.

In an elastic collision, the objects separate after impact and don’t lose any of their kinetic energy. Kinetic energy is the energy of motion and is covered in detail elsewhere. The law of conservation of momentum is very useful here, and it can be used whenever the net external force on a system is zero. The below figure shows an elastic collision where momentum is conserved

Perfectly elastic collisions can happen only with subatomic particles. Everyday observable examples of perfectly elastic collisions don’t exist—some kinetic energy is always lost, as it is converted into heat transfer due to friction. However, collisions between everyday objects are almost perfectly elastic when they occur with objects and surfaces that are nearly frictionless, such as with two steel blocks on ice.

Now, to solve problems involving one-dimensional elastic collisions between two objects, we can use the equation for conservation of momentum. First, the equation for conservation of momentum for two objects in a one-dimensional collision is

Substituting the definition of momentum p = mv for each initial and final momentum, we get

where the primes (‘) indicate values after the collision; In some texts, you may see i for initial (before collision) and f for final (after collision). The equation assumes that the mass of each object does not change during the collision

Now, let us turn to the second type of collision. An inelastic collision is one in which objects stick together after impact, and kinetic energy is not conserved. This lack of conservation means that the forces between colliding objects may convert kinetic energy to other forms of energy, such as potential energy or thermal energy. The concepts of energy are discussed more thoroughly elsewhere. For inelastic collisions, kinetic energy may be lost in the form of heat. Below figure shows an example of an inelastic collision. Two objects that have equal masses head toward each other at equal speeds and then stick together. The two objects come to rest after sticking together, conserving momentum but not kinetic energy after they collide. Some of the energy of motion gets converted to thermal energy, or heat.

The figure shows A one-dimensional inelastic collision between two objects. Momentum is conserved, but kinetic energy is not conserved. (a) Two objects of equal mass initially head directly toward each other at the same speed. (b) The objects stick together, creating a perfectly inelastic collision. In the case shown in this figure, the combined objects stop; This is not true for all inelastic collisions.

Since the two objects stick together after colliding, they move together at the same speed. This lets us simplify the conservation of momentum equation from

to

for inelastic collisions, where v′ is the final velocity for both objects as they are stuck together, either in motion or at rest.

But what about collisions, such as those between billiard balls, in which objects scatter to the side? These are two-dimensional collisions, and just as we did with two- dimensional forces, we will solve these problems by first choosing a coordinate system and separating the motion into its x and y components.

We start by assuming that Fnet = 0, so that momentum p is conserved. The simplest collision is one in which one of the particles is initially at rest. The best choice for a coordinate system is one with an axis parallel to the velocity of the incoming particle, as shown in below figure . Because momentum is conserved, the components of momentum along the x– and y-axes, displayed as px and py, will also be conserved. With the chosen coordinate system, py is initially zero and px is the momentum of the incoming particle.

Now, we will take the conservation of momentum equation, p1 + p2 = p1 + p2 and break it into its x and y components. Along the x-axis, the equation for conservation of momentum is

In terms of masses and velocities, this equation is

But because particle 2 is initially at rest, this equation becomes

The components of the velocities along the x-axis have the form v cos θ . Because particle 1 initially moves along the x-axis, we find v1x = v1. Conservation of momentum along the x-axis gives the equation

where θ1 and θ2 are as shown in the above Figure

Along the y-axis, the equation for conservation of momentum is

or

But v1y is zero, because particle 1 initially moves along the x-axis. Because particle 2 is initially at rest, v2y is also zero. The equation for conservation of momentum along the y-axis becomes

The components of the velocities along the y-axis have the form v sin θ  . Therefore, conservation of momentum along the y-axis gives the following equation:

(You can access this textbook for free in web view or PDF through OpenStax.org, and for a low cost in print)

Take Quiz

1) Describe the transformation between forms of mechanical energy that is happening to a falling skydiver before his parachute opens.
a. Kinetic energy is being transformed into potential energy.
b. Potential energy is being transformed into kinetic energy.
c. Work is being transformed into kinetic energy.
d. Kinetic energy is being transformed into work.

ANSWER

b. Potential energy is being transformed into kinetic energy.

2) True or false—If a rock is thrown into the air, the increase in the height would increase the rock’s kinetic energy, and then the increase in the velocity as it falls to the ground would increase its potential energy.
a. True
b. False

ANSWER

b. False

3) A 0.2 kg apple on an apple tree has a potential energy of 10 J. It falls to the ground, converting all of its PE to kinetic energy. What is the velocity of the apple just before it hits the ground?
a. 0 m/s
b. 2 m/s
c. 10 m/s
d. 50 m/s

ANSWER

c. 10 m/s

4) Is it possible for the sum of kinetic energy and potential energy of an object to change without work having been done on the object? Explain.
A) No, because the work-energy theorem states that work done on an object is equal to the change in kinetic energy, and change in KE requires a change in velocity. It is assumed that mass is constant.
b) No, because the work-energy theorem states that work done on an object is equal to the sum of kinetic energy, and the change in KE requires a change in displacement. It is assumed that mass is constant.
C) Yes, because the work-energy theorem states that work done on an object is equal to the change in kinetic energy, and change in KE requires a change in velocity. It is assumed that mass is constant.
D)Yes, because the work-energy theorem states that work done on an object is equal to the sum of kinetic energy, and the change in KE requires a change in displacement. It is assumed that mass is constant

ANSWER

D) Yes, because the work-energy theorem states that work done on an object is equal to the sum of kinetic energy, and the change in KE requires a change in displacement. It is assumed that mass is constant

5) The starting line of a cross country foot race is at the bottom of a hill. Which form(s) of mechanical energy of the runners will change when the starting gun is fired?
a. Kinetic energy only
b. Potential energy only
c. Both kinetic and potential energy
d. Neither kinetic nor potential energy

ANSWER

c. Both kinetic and potential energy

6) True or false—A cyclist coasts down one hill and up another hill until she comes to a stop. The point at which the bicycle stops is lower than the point at which it started coasting because part of the original potential energy has been converted to a quantity of heat and this makes the tires of the bicycle warm.
a. True
b. False

ANSWER

b. False

7) A boy pushes his little sister on a sled. The sled accelerates from 0 to 3.2 m/s . If the combined mass of his sister and the sled is 40.0 kg and 18 W of power were generated, how long did the boy push the sled?
a. 205 s
b. 128 s
c. 23 s
d. 11 s

ANSWER

c. 23 s

8) What is the kinetic energy of a bullet traveling at a velocity of ?
a. 3.5 L
b. 7J
c. 2.45 x 103 J
d. 2.45 x 105 J

ANSWER

c. 2.45 x 103 J

9) A marble rolling across a flat, hard surface at rolls up a ramp. Assuming that and no energy is lost to friction, what will be the vertical height of the marble when it comes to a stop before rolling back down? Ignore effects due to the rotational kinetic energy.
a. 0.1 m
b. 0.2 m
c. 0.4 m
d. 2 m

ANSWER

b. 0.2 m

10) The work–energy theorem states that the change in the kinetic energy of an object is equal to what?
a. The work done on the object
b. The force applied to the object
c. The loss of the object’s potential energy
d. The object’s total mechanical energy minus its kinetic energy

ANSWER

a. The work done on the object

11) A runner at the start of a race generates 250 W of power as he accelerates to 5 m/s . If the runner has a mass of 60 kg, how long did it take him to reach that speed?
a. 0.33 s
b. 0.83 s
c. 1.2 s
d. 3.0 s

ANSWER

d. 3.0 s

12) A car’s engine generates 100,000 W of power as it exerts a force of 10,000 N. How long does it take the car to travel 100 m?
a. 0.001 s
b. 0.01 s
c. 10 s
d. 1,000 s

ANSWER

c. 10 s

13) Why is this expression for kinetic energy incorrect?
KE = (m) (v) 2
a. The constant is missing.
b. The term should not be squared.
c. The expression should be divided by 2 .
d. The energy lost to friction has not been subtracted

ANSWER

c. The expression should be divided by 2 .

14) What is the kinetic energy of a 10kg object moving at 2.0 m/s ?
a. 10 J
b. 20 J
c. 40 J
d. 100 J

ANSWER

b. 20 J

15) A boulder rolls from the top of a mountain, travels across a valley below, and rolls part way up the ridge on the opposite side. Describe all the energy transformations taking place during these events and identify when they happen.
a. As the boulder rolls down the mountainside, KE is converted to PE. As the boulder rolls up the opposite slope, PE is converted to KE. The boulder rolls only partway up the ridge because some of the PE has been converted to thermal energy due to friction.
b. As the boulder rolls down the mountainside, KE is converted to PE. As the boulder rolls up the opposite slope, KE is converted to PE. The boulder rolls only partway up the ridge because some of the PE has been converted to thermal energy due to friction.
c. As the boulder rolls down the mountainside, PE is converted to KE. As the boulder rolls up the opposite slope, PE is converted to KE. The boulder rolls only partway up the ridge because some of the PE has been converted to thermal energy due to friction.
d. As the boulder rolls down the mountainside, PE is converted to KE. As the boulder rolls up the opposite slope, KE is converted to PE. The boulder rolls only partway up the ridge because some of the PE has been converted to thermal energy due to friction.

ANSWER

d. As the boulder rolls down the mountainside, PE is converted to KE. As the boulder rolls up the opposite slope, KE is converted to PE. The boulder rolls only partway up the ridge because some of the PE has been converted to thermal energy due to friction.

Read more
  • 1
  • 2
Back to Top
error: Content is protected !!
View Chapters