Viscous 9

In the previous chapter on fluids, we introduced the basic ideas of , flow, the application of conservation of mass and of energy in the form of the continuity equa- tion and of Bernoulli’s equation, respectively, as well as . Throughout those discussions we restricted ourselves to ideal fluids, those that do not exhibit any frictional properties. Often these can be neglected and the results of the previous chapter applied without any modifications whatsoever. Clearly mass is conserved even in the presence of viscous frictional forces and so the continuity equation is a very general result. Real fluids, however, do not conserve mechanical energy, but over time will lose some of this well-ordered energy to heat through frictional losses. In this chapter we consider such behavior, known as , first in the case of simple fluids such as water. We study the effects of viscosity on the motion of simple fluids and on the motion of suspended bodies, such as macromolecules, in these fluids, with special attention to flow in a cylinder, the most important geometry of flow in biology. The complex nature of blood as a fluid is studied next leading into a description and physics perspective of the human circulatory system. We conclude the chapter with a discussion of and capillarity, two important surface phenomena in fluids. In Chapter 13 we return to the general notion of the loss of well-ordered energy to heat in the context of thermodynamics.

1. VISCOSITY OF SIMPLE FLUIDS

Real fluids are viscous, having internal attractive forces between the molecules so that any relative motion of molecules results in frictional, or drag, forces. The work done by these drag forces, in turn, results in a loss of mechanical energy due to slight heat- ing. We can think of viscosity as a measure of the resistance of a liquid to flowing, so that liquids such as paint or maple syrup have much higher than water. A quantitative definition of viscosity can be introduced using the example of laminar flow of a liquid between two parallel plates (Figure 9.1), the lower one fixed and the upper one pulled by an external force to move with a constant velocity v parallel to the surface of the plate. Clearly in the absence of drag forces the constant external force would lead to uniform acceleration of the top plate, but due to the drag forces the top plate quickly reaches a steady-state constant velocity. Because the liquid is viscous, it tends to stick to the surfaces of the plates, forming a boundary layer. Therefore the liq- uid layer at the fixed plate is at rest, whereas the liquid layer at the top plate moves with velocity v. For laminar flow, the velocity of the liquid varies linearly in the trans- verse direction (y-direction in Figure 9.1) from 0 to v over the separation distance between the plates of area A. Planar layers of fluid slide over one another. Viscosity can be defined through the relation between the shear , or force per unit area F/A, needed to keep the upper plate moving with a constant velocity

J. Newman, Physics of the Life Sciences, DOI: 10.1007/978-0-387-77259-2_9, © Springer Science+Business Media, LLC 2008

V ISCOSITY OF S IMPLE F LUIDS 231 y and the rate of variation of the velocity between the plates, ⌬v/⌬y (known as the rate of strain), F ¢v ϭ h , (9.1) v A ¢y where ␩ is the viscosity of the liquid. Contrast this with the stressÐstrain relation discussed in Chapter 3 for solids where the strain ⌬x/⌬y appeared on the right-hand side and not the rate of strain, appropriate here for fluids. Strain and rate of strain are connected in the usual way because the time rate of change of strain is given by (⌬x/⌬y)/⌬t ϭ (⌬x/⌬t)/⌬y ϭ⌬v/⌬y. The SI unit for viscosity is the Pa-s, but another commonly used unit is the poise ϭ FIGURE 9.1 A fluid sandwiched (P; 1 P 10 Pa-s). Table 9.1 lists viscosities of water and blood. Equation between two plates with the bot- (9.1) can be taken as the definition of viscosity, originally due to Sir Isaac tom plate fixed and the top plate Newton. Fluids that obey this relation are said to be Newtonian fluids. The proportion- moving at a constant velocity v. ality of the shear stress and rate of strain usually holds only at lower strain rates. Water and salt solutions are Newtonian, whereas blood, whose behavior does not follow Equation (9.1), is said to be a non-Newtonian fluid and is discussed in the next section.

Table 9.1 Viscosities of Water and Blood

Fluid Temperature Viscosity (10Ϫ3 Pa-s)

Water 0 1.8 20 1.0 37 0.7 Whole blooda 37 4.0 Blood plasma 37 1.5

a Varies greatly with hematocrit, or red blood cell content.

When a solid is put under shear stress, with an external force applied in a partic- ular direction, it deforms and, for small stresses F/A, the strain, or response of the solid, is proportional to the stress. Once the stress is removed, the solid returns to its original shape (unless it has some plasticity, in which case it may flow). In a Newtonian liquid, however, a constant applied shear stress results in a constant rate of strain (Equation (9.1)) rather than constant strain. The larger the rate of strain, meaning the more abruptly the velocity changes with transverse distance, the greater the viscous force, and in turn, the greater the applied shear stress needed to keep the top plate moving at the same constant velocity. At higher shear stress there are devi- ations from this relation, and at still higher stress, turbulence will occur.

Example 9.1 A sheet of plywood is covered with a 1 mm thick layer of tile adhesive and a square piece of ceramic tile measuring 30 cm on a side is placed on it. If a force of 10 N is applied parallel to the surface, find the velocity with which the tile slides. Assume laminar Newtonian flow and use a viscosity of 50 Pa-s for the adhesive.

Solution: We first calculate the stress as F/A ϭ 10/(.3)2 ϭ 110 N/m2. Dividing this stress by the adhesive viscosity, the rate of strain is found to be 2.2 sϪ1, so that the velocity of the tile is given as ¢v v ϭ y ϭ (2.2 sϪ1)(1 mm) ϭ 2.2 mm/s. ¢y

FIGURE 9.2 Laminar capillary flow showing a concentric layer of fluid that flows at the same velocity The capillary tube is a very common geometry for fluid flow in biology. It is along the length of the tube. relevant for blood flow, for example, as well as for viscometry, the methodology of

232 V ISCOUS F LUIDS viscosity measurement. When a liquid flows through a tube 1 without obstacles, the flow at low velocities is laminar with lay- 0.75 ers of liquid in concentric cylinders (Figure 9.2). The outermost 0.5 layer is the boundary layer that remains at rest and the fastest v/vo flowing liquid lies at the center of the tube. The actual velocity 0.25 profile across the tube is parabolic as indicated in Figure 9.3. 0 The velocity varies across the capillary tube; thus in order to 0 0.2 0.4 0.6 0.8 1 find the volume flow rate, Q (ϭ vA when the velocity was r/ro assumed uniform in the absence of viscosity), an average must FIGURE 9.3 The velocity profile be calculated across the cross-sectional area. This was first done in 1835 by across a capillary tube of radius ro. Poiseuille, a French physician interested in blood flow (the viscosity unit poise is taken from his name), who found

pPr4 Q ϭ , (9.2) 8hL where P/L is the applied pressure per unit length of the tube and r is the tube radius. Equation (9.2) is known as Poiseuille’s law. If we rewrite this equation in the form

8hL ¢P ϭ Q, (9.3) a pr4 b where we write ⌬P as the pressure difference across the tube of length L, then we can interpret the equation as follows. For a given ⌬P across the tube, the resulting flow Q depends on the resistive term in parentheses. The larger this term, the slower the flow rate is for a given applied pressure. With a constant resistive term (fixed tube length, radius, and fluid viscosity), the greater the pressure difference acting on the liquid, the greater is the expected flow rate. A longer tube or larger viscosity provides a greater resistance to flow as might be intuitively expected. The very strong dependence of the resistive term on tube radius rϪ4 is surprising and extremely significant in controlling the flow rate of a liquid in a capillary tube. The resistance to fluid flow increases dramatically as the tube radius gets smaller. This can lead to important effects in the flow of blood in arteries because a partially clogged artery will require a much higher pressure differential to supply the same fluid flow rate.

Example 9.2 In giving a transfusion, blood drips from a sealed storage bag with a 1 m pressure head through capillary tubing of 2 mm inside diameter, passing through a hypodermic needle that is 4 cm long and has an inside diameter of 0.5 mm. If the blood pressure within the vein into which the blood is being trans- fused is at a gauge pressure of 18 torr, find how long it will take to give the patient 1 L of blood. How long will it take if the inside diameter of the needle is only 0.4 mm?

Solution: Since the flow rate depends so strongly on the radius of the capillary, the most resistance to flow will occur within the hypodermic needle and rela- tively little within the delivery tubing. We can therefore apply Equation (9.2) using the radius and length of the needle, ignoring the dimensions of the tubing. For the net driving pressure across the column of blood up to the vein we use a value of P ϭ (␳gh Ϫ 18 torr) ϭ (␳gh Ϫ 2400 N/m2), where the density of blood (Continued)

V ISCOSITY OF S IMPLE F LUIDS 233 ϭ 5 2 ϭ is found in Table 9.1 and we have used the conversion from Patm 10 N/m 760 torr. We find a flow rate of

pPr4 p[(1.06 ϫ 103)(9.8)(1) Ϫ 2400](0.00025)4 Q ϭ ϭ 8hL 8(4 ϫ 10Ϫ3)(0.04) ϭ 7.7ϫ10Ϫ8 m3/s ϭ 0.077 cm3/s.

With this flow rate, each cm3 of blood will take 13 s to flow into the vein, so that it will take a total time of 3.6 h for a liter of blood to be transfused. If the r value is 0.2 mm then the flow rate will decrease by the factor (2/2.5)4 ϭ 0.41 and so it will take 3.6/0.41 ϭ 8.8 h. We see that a decrease in the radius by a fac- tor of only 0.8 increases the time required by almost 2.5 times, pointing out the very strong dependence on r.

Capillary viscometers make use of Poiseuille’s law to measure the relative vis- cosity of liquids or solutions. They consist of a fine capillary tube in which a liquid is placed and measurements made of the time for a fixed volume of liquid to flow through the tube (Figure 9.4). Because the pressure P is equal to ␳gh, where h is the height of liquid in the tube, we find from Equation (9.2) that for a given capillary tube Q ϰ ␳/␩, where the other parameters are independent of the liquid properties. Q is a flow rate and therefore Q ϰ 1/t, where t is the time for a fixed fluid volume to flow through the capillary, so we have that

r 1 r h rrt. h t or FIGURE 9.4 Capillary viscometer used to measure solvent viscosity From measurements of the efflux times of the same volume of by a timing measurement. an unknown fluid and a standard fluid, we can take the ratio to write that

r t ϭ unknown unknown hunknown hstnd . (9.4) rstnd tstnd

If the densities and standard viscosity are known, the viscos- ity of the unknown liquid can be determined from simple timing measurements. Results from such measurements can give accurate viscosity values for pure liquids or for solvents (typically solutions of small dissolved ions). Thus far we have considered the flow of pure viscous fluids at low stress. At higher stress turbulence occurs and the flow profile in a capillary is much different than for laminar flow (Figure 9.5). In such turbulent flow there is a much greater effec- tive internal friction due to vortices, and also the strain rate ⌬v/⌬r near the walls is much greater (note the more rapid velocity change near the boundary layer on the tube wall in Figure 9.5). What happens when a flowing viscous fluid meets an obsta- cle, perhaps a biological macromolecule? We have already briefly considered this question in our discussion of motion in a fluid in Chapter 3, Section 2. There we introduced the dimen- sionless Reynolds number t , defined as

Lrv t ϭ , h (9.5)

234 V ISCOUS F LUIDS where L is the characteristic size of the object. Imagine a sphere of radius v v r held fixed within a flowing fluid. Under laminar flow conditions, with t on the order of 1 or less, the fluid will flow around the sphere in a sym- metric pattern as shown in Figure 9.6 (top). There is a frictional force that acts on the sphere given by Stokes’ law,

ϭ - Ff 6phrv. (9.6) laminar turbulent FIGURE 9.5 Velocity profiles for The frictional force in this case varies linearly with both the fluid laminar and turbulent capillary flow. velocity and viscosity as well as with the size of the sphere. As the fluid velocity is Note that the profile of the fluid in increased, the flow pattern will become more complex and asymmetric, and the fric- the tube is not shown here, but tional force will become dependent on the square of the fluid velocity, as already dis- rather how the velocity varies across the capillary. cussed (see Equation (3.5)). The fluid velocity downstream from the sphere is decreased as the Reynolds number is increased, and at a certain point the flow becomes unsteady with “eddies,” or vortices, forming in the downstream region known as the wake of the object (Figure 9.6, middle); at even higher Reynolds num- bers the flow becomes fully turbulent (Figure 9.6, bottom). By careful design of the shape of an object, the frictional forces can be reduced. Engineered streamlined designs have led to improved aerodynamic performance of cars and airplanes. In the world of animals, evolutionary design has also resulted in streamlined shapes partic- ularly for many aquatic or flying animals. The problem of determining the viscosity of a suspension of objects is a very complex one. When more than one object is present in a fluid, the wake produced by one object can interact with the other objects through what are termed hydrodynamic interactions. In 1906, Einstein solved the problem of determining the viscosity of a ␩ suspension of identical spherical particles s and found

ϭ ϩ £ hs ho (1 2.5 ), (9.7)

␩ ⌽ where o is the solvent viscosity and is the (dimensionless) volume fraction occupied by the spheres. Note that this result does not depend explicitly on the particle radius. The larger the sphere is, the smaller the number of them required to occupy the same volume fraction and hence have the same solution viscosity. For particles of other shapes the factor 2.5 is replaced by a shape-dependent numer- ical factor.

Example 9.3 Find the viscosity of a 100 ␮M aqueous solution of a small spher- ical protein with radius 5 nm and molecular weight 40,000 at 20¡C. This might be a solution of globular actin protein.

Solution: To proceed from Equation (9.7), we need to calculate the volume fraction occupied by the protein. Each protein molecule occupies a volume of Ϫ 4Ð ␲ r3 ϭ ϫ 25 3 N 3 5.2 10 m and each protein molecule has a mass of 40,000/ A, ϫ Ϫ20 ϭ ϫ Ϫ23 where NA is Avogadro’s number, or a mass of 6.6 10 g 6.6 10 kg. A 100 ␮M solution has a density of 40,000 g/mol ϫ 10Ϫ4 mol/L ϭ 4 g/1000 cm3 ϭ 4 ϫ 10Ϫ3 kg/10Ϫ3 m3 ϭ 4 kg/m3, using 1 cm3 ϭ 10Ϫ6 m3. We can then compute that in every unit volume (1 m3) there are 4/(6.6 ϫ 10Ϫ23) ϭ 6.1 ϫ 1022 mole- cules occupying a volume of (6.1 ϫ 2022)(5.2 ϫ 10Ϫ25) ϭ 0.03 m3. Thus the ␩ ϭ Ϫ3 volume fraction is 0.03 and the viscosity is then found to be (using o 10 Pa-s for water) ␩ ϭ [1 ϩ (2.5)(0.03)] ϫ 10Ϫ3 Pa-s ϭ 1.075 ϫ 10Ϫ3 Pa-s, a 7.5% increase over pure water.

V ISCOSITY OF S IMPLE F LUIDS 235 Viscosities of suspensions or solutions of macromolecules can be measured using capillary viscometers, just as for pure fluids, if the par- ticles or macromolecules are small, so that they are not oriented by the flow in a capillary, and if a sufficient volume of material is available (typically 0.1 L). Other designs for viscometers have been developed to use smaller volumes and to extrapolate to zero shear rate in order to avoid orienting asymmetric particles. When a DNA molecule is stretched by hydrodynamic forces during flow, it responds somewhat like a stretched rubber band, storing energy like a spring that can be recovered when the flow stops. Elastic proper- ties of DNA and many other biomolecules seem to be very important in their functioning. Solutions of DNA and other fiberlike molecules (filamentous proteins and other elongated (bio)polymers) exhibit vis- coelasticity, having both a measurable viscosity, or energy loss mecha- nism, as well as elastic storage of energy. One method by which such solutions can be studied involves more sophisticated viscometers, called rheometers (after rheology, the study of viscoelasticity), in which both the elasticity and viscosity are simultaneously measured to give infor- mation about the structure and functioning of these macromolecules. FIGURE 9.6 Flow patterns around a sphere at increasing Reynolds numbers. 2. BLOOD AND OTHER COMPLEX FLUIDS

The term “complex fluid” is usually used for a non-Newtonian fluid, meaning that the shear stress and rate of strain are not simply proportional as they are in Equation (9.1). Most biological fluids are complex, including blood. Even simple suspensions of asym- metric macromolecules are non-Newtonian due to orientation effects at higher strain rates: large transverse variations in velocity create torques on such molecules tending to align them in the flowing fluid, just as a stick aligns itself with the flow in a fast-moving stream. Other complex biological “fluids” include cellular cytoplasm, which has viscoelastic FIGURE 9.7 (top) The relation properties, and biological membranes, having two-dimensional fluidlike properties briefly between viscosity and shear rate discussed in Section 6 of Chapter 7. In this section we consider the composition and prop- for whole blood (with the hemat- ocrit shown) and blood plasma; erties of blood as perhaps the most interesting example of a complex fluid. (bottom) the viscosity of whole Human blood makes up about 1/13 of the total body blood versus hematocrit. mass and amounts to 5Ð6 L in the average adult male. 100000 When blood is centrifuged it separates into two portions. Plasma is the fluid component of blood and is composed 1000 by weight of about 92% water, 7% protein, and small 90% amounts of organic and inorganic molecules as well as 10 45% dissolved gases. It behaves as a Newtonian viscous fluid

viscosity (cP) 0% with a viscosity about 20% higher than that of water. The 0.1 second phase that spins down in a centrifuge consists of 0.01 0.1 1 10 100 1000 cells, primarily red blood cells that make up over 50% of shear rate (1/s) the volume of blood. Red cells, or erythrocytes, contain hemoglobin and carry oxygen throughout the body. There are also much smaller numbers of white blood cells and platelets in blood. The white cells, or leuko- cytes, come in five varieties and are capable of amoeboid motion and one variety, the neutrophils, can migrate out of small blood vessels and play a role in fighting infec- tions by engulfing bacteria throughout the body in a process called phagocytosis. Platelets are small cells that are involved in blood clotting. All of these cells have finite life spans ranging from one or two days to several months and are replenished by the bone marrow. Figure 9.7 (top) shows data for the viscosity of whole blood at three different hematocrits (the percent

236 V ISCOUS F LUIDS FIGURE 9.8 Red blood cells (left: showing biconcave shape and right: red cells aggregating to form stacked cells, or rouleaux). of blood volume occupied by cells) as a function of the shear rate. The non- Newtonian feature is the variation of viscosity by large factors (note the log scales) with shear rate for blood with cells present. Blood plasma is a Newtonian fluid because its viscosity is independent of shear rate (lower curve). The red blood cells normally constitute about 50% of the blood volume, therefore it is clear that the non- Newtonian rheological properties of blood are primarily due to the red cells. In the bottom half of Figure 9.7 the low-shear viscosity of whole blood is shown as a func- tion of the hematocrit. The strong dependence on the red cell content is also indica- tive of the large impact of the red cells on the rheological properties of blood. Red blood cells are disks that are biconcave (thinner in the middle than at the edges), are about 8 ␮m in diameter and have a tendency to stack together like coins, into aggre- gates called rouleaux (Figure 9.8). The extent of aggregation is strongly dependent on the shear rate; the aggregates will break up as the shear rate is increased, qualita- tively explaining the decrease in viscosity at increasing shear rates shown in the top of Figure 9.7. Blood is remarkably fluid. A 50% (by volume) suspension of small rigid spheres will be a solid, unable to flow at all, whereas blood is extremely fluid even at elevated hematocrits (Figure 9.9). This fluidity is due to the special properties of FIGURE 9.9 Viscosity, relative to the red blood cells, particularly their membrane elastic properties and shape, which water, for human blood (lower permit tremendous deformation of the red cells to allow flow. In many small blood curve) and a suspension of rigid vessels, the capillary diameters are on the order of the red cell diameter or even plastic spheres (upper curve) as a smaller and without great flexibility of the red cells, flow would be blocked. function of the volume fraction Diseased red cells, such as deformed cells in sickle cell anemia that lose their elas- occupied by particles. At volume fractions of about 50% or higher a tic properties, will clog small blood vessels. In the next suspension of plastic spheres section we take up the human circulatory system, behaves as a solid. including the heart, and expand on the flow properties 1000 of blood.

100 3. THE HUMAN CIRCULATORY SYSTEM 10 In Western culture, the concept of blood circulation was relative viscosity established surprisingly late, in the 1600s, by William Harvey. The human circulatory system consists of a 1 pump (the heart) and a complex branched distribution of 0 0.2 0.4 0.6 0.8 1 “smart” delivery tubes that carry oxygen and nutrients volume fraction

T HE H UMAN C IRCULATORY S YSTEM 237 FIGURE 9.10 (a) Schematic diagram of the heart and the flow of blood. Red indicates oxygenated blood and blue deoxygenated in this lungs simplified scheme. (b) Schematic Left Atrium open model of the heart. Right Atrium RA LA Left Ventricle aorta RV LV

Right Ventricle

capillary bed

to, and remove waste products from, the body. Each side of the heart receives blood at low pressure and pumps this blood out at high pressure. In schematic form, shown in Figure 9.10a, oxygenated blood is pumped out of the left ventricle of the heart, through the aortic valve and the aorta to a branched network of arteries, smaller arte- rioles, and finally to the capillary beds throughout the body in which the exchange of gases and dissolved molecules with the body tissue occurs. Blood is collected from the capillary beds by the venules, which feed into the veins, all of which merge with either the superior (from above the heart) or inferior vena cava (from below the heart) or the coronary sinus (from the blood supply for the heart muscle itself) to return the blood to the right atrium. Thus the left ventricle and the right atrium of the heart together form the outlet and inlet of a pump that supplies nearly the entire body with blood. A second parallel pump in the heart sends blood that has arrived from the right atrium through the tricuspid valve to the right ventricle, through the pulmonary arter- ies to the lungs where an exchange of gases occurs. The reoxygenated blood returns to the left atrium through the pulmonary veins where it enters the left ventricle through the bicuspid (mitral) valve to complete its cycle of flow. Thus, the right ven- tricle and left atrium are the outlet and inlet for a second pump of the heart. In the healthy mammalian heart the chambers of the left side of the heart are completely separated from those of the right after birth, and there is no mixing of oxygenated and deoxygenated blood in the heart. Despite this separation, the two sides are part of a single anatomical organ, and the heartbeat is coordinated by a single clump of cells, the pacemaker region. A schematic of the heart is shown in Figure 9.10b. In the rest of this section we consider several aspects of the circulatory system that relate to our previous discussions in this chapter. We return later (Chapter 15) to consider the elec- trical aspects of the heart, including the electrocardiogram (EKG). The heart, about the size of an adult fist, pumps about 80 cm3 of blood in each FIGURE 9.11 A single cardiac of the 70 beats/minute in a typical resting adult, so that about 5.5 L of blood are cycle showing the left ventricular pumped throughout the body each minute. Because the total volume of blood in an pressure (bold) and aortic pressure adult is 5Ð6 L, we conclude that it takes just about a minute for blood to make a com- (dashed) as functions of time. Also indicated are the times at which plete loop through the circulatory system. The total volume of blood is actually in valves open and close, when heart dynamic equilibrium because fluids leave the blood vessels to exchange with tissue sounds are most clear, and the and to be filtered in the kidneys (discussed in Chapter 12). Figure 9.11 period of systole and diastole. shows some events during a single cardiac cycle, divided into the sys- tole, or contraction, phase and the diastole, or relaxation, phase. Note 120 close open that the left and right ventricles contract together, as do the atriums. During systole, the ventricular pressure rises rapidly, after closure of the 80 systole pressure tricuspid or mitral valve, as the blood volume in the ventricle increases. (Torr) diastole When the aortic valve opens, the aortic pressure rises from its resting sounds value of about 80 mm Hg to about 120 mm Hg. It is this pressure that is measured with a sphygmomanometer. Figure 9.11 also shows the times at which valves open and close and time those at which the heart sounds are most clear. The pulmonary artery

238 V ISCOUS F LUIDS pressure rises during the contraction of the right ventricle, but to a lesser extent than in the aorta; the peak pressure in the aorta is about six times that of the pulmonary artery. The greater pressure generated by the left ventricle is a result of thicker layers of muscle surrounding it compared to the right ventricle. in the atria are close to zero and only fluctuate a little during the cardiac cycle. As an example of applying some of the fluid dynamics we have learned, we can make an estimate of the power developed by the heart in pumping blood, where power here is the time rate of transfer of energy to the blood. The heart supplies both the pressure and kinetic energy of the blood as it leaves the heart and enters the aorta. If we multiply Bernoulli’s equation for constant height (see Equation (8.12)) by the volume flow rate Q, we obtain an expression for the power supplied by the heart as

1 Power ϭ P Q ϩ rv2Q, ave 2 ␳ where Pave is the mean blood pressure in the aorta, is the density of blood, and v is the average blood velocity in the aorta. If we take the mean aortic pressure to be 100 mm Hg when a person is at rest, then the PQ contribution of the left ventricle in one heartbeat is simply the product of the mean pressure and the volume change, 80 cm3, resulting in a value of (100 mm Hg) ϫ (1.01 ϫ 105 Pa/760 mm Hg) ϫ (80 ϫ 10Ϫ6 m3) ϭ 1.06 J/heartbeat. Assuming 70 heartbeats per minute (or 1.2 beats/s) this translates into an average power of about 1.3 W. Because the pressure in the right ventricle is about 1/6 that of the left ventricle and the volume flow rate is the same, the PQ power contribu- tion of the right ventricle is an additional 0.2 W. The kinetic energy term contributes a small additional amount of about 0.3 W when a person is at rest, so that the total power supplied by the heart is about 1.8 W. To find this kinetic energy contribution we use the fact that Q ϭ Av, or v ϭ Q/A, so that the kinetic energy term is 1/2rv2Q ϭ 1/2r(Q3/A2), proportional to Q3. When someone engages in very strenuous exercise, the flow rate of blood can reach 35 L/min (nearly 7 times the resting rate). In this case, assuming the mean pressure does not change significantly, the PQ power increases by a factor of 7 to about 10 W, and the kinetic energy power delivered to the blood rises dramatically to (0.3 W) (73) ഡ 100 W, because of its third-order dependence on Q. Where does this power go? Just ask someone doing exercise and they will tell you how hot they get and the amount of sweating they do in an attempt to cool off. The key to the heart’s success in maintaining pressure differentials in order to drive blood throughout the body is the four heart valves. Heart valves are crucial for the proper functioning of the heart and a number of heart diseases are traceable to defective valves. Perhaps surprisingly, the heart valves (and those of the veins men- tioned below) are not controlled actively, but open and close passively in response to hydrodynamic forces. Consider the mitral valve, located between the left atrium and ventricle, shown schematically in Figure 9.12. In the diastole, when the pressure in the atrium exceeds that in the ventricle, the two thin membranes of the valve are pushed open and blood enters the ventricle. As the blood pours into the ventricle it strikes the ventricular walls and the flow breaks up into eddies, or vortices, that pro- vide a back-pressure on the valve membranes, forcing them closed when the ventric- ular pressure exceeds the atrial pressure. A set of small muscles prevents the membranes from opening to allow backflow into the atrium; when properly func- tioning, the mitral valves prevent any blood from re-entering the atrium. Some types of heart murmurs are due to malfunctioning heart valves that allow backflow, pro- ducing characteristic sounds. Heart valves used in an artificial heart also make use of the same principles to provide passive control, rather than direct active control of the opening and closing of valves. The cyclic variation in the aortic (and pulmonary arterial) pressure is the driving force producing blood flow throughout the body (and the lungs). We have seen from the continuity equation that the flow velocity in a tube is inversely

T HE H UMAN C IRCULATORY S YSTEM 239 FIGURE 9.12 (left two panels) The closure of the mitral valve by the backflow of blood in the left ventricle. (right panel) Schematic of the heart valves.

proportional to the cross-sectional area, in order to conserve mass. In the circula- tory system, one large artery (typical inside diameter 1 cm) divides into many arterioles (typical diameter 5 ␮m), each of which divides into many capillaries (typical diameters 0.6 ␮m). The capillaries have the smallest diameter therefore we might expect blood to flow fastest in these vessels. Although the capillaries do have the smallest cross-section, the total cross-sectional area of the estimated five billion capillaries is about five times that of the arterioles. Blood velocity in a cap- illary is therefore slower than in any other blood vessel, only about 0.07 cm/s. Capillary diameters are comparable to the dimensions of a red blood cell and so the flow of blood through capillaries, known as bolus flow, is quite special. As shown in Figure 9.13, to promote the flow of the red cells and the exchange of gases and chemicals across the vessel walls, the elastic red cells trap blood plasma between themselves that flows in eddies. In some regions, blood flow from the arterioles can bypass the capillaries and flow directly into venules through an arteriovenous (AV) shunt. These shunts are able to regulate the flow of blood in order to control, for example, the extent of body cooling through blood flow in the skin. During exercise, as metabolism is increased, or when the external temperature is high, excess heat must be removed by evaporation and the capillaries near the skin surface are dilated by decreasing the AV shunt flow. Similarly in cold weather, the AV shunt is opened to decrease blood flow near the skin surface in order to reduce heat loss from the body. Another control mechanism outside the heart is vasoconstriction, a reflex process of reducing the diameters of blood vessels to increase flow rates in the case of blood loss or shock. Blood flow in the larger arteries is known as pulsatile flow. As the ventricles pump blood into the major arteries, the blood cannot flow into the capillaries fast enough and so the arteries swell in diameter because the walls are elastic. As the FIGURE 9.13 Bolus flow of red pressure in the artery drops during diastole, the energy stored in the elastic vessel blood cells moving to the left in a walls tends to smooth out pressure variations and this becomes more and more the capillary, showing the eddy flow between cells. case farther downstream from the aorta. This same elastic expansion of blood vessels can be felt as the pulse measured at one’s wrist. By the time the blood leaves the capillary bed, the pressure in the veins is quite low. To help the return flow, larger veins, particularly in the limbs, have one-way valves along them. Excess fluid pressure in the feet, due to the extra pressure head, can sometimes result in fluid buildup and swelling (edema), especially without movement of the feet in order to promote blood flow in the venules and veins.

240 V ISCOUS F LUIDS 4. SURFACE TENSION AND CAPILLARITY Δx

The surface of a fluid represents a boundary that exhibits many special properties wor- thy of our attention. A thin layer of surface fluid in contact with air feels an excess attractive intermolecular force over the local interactions within the bulk fluid. The net F force pulling the surface layer into the bulk fluid gives rise to a slightly greater den- w sity near the surface. Molecules that move into the surface layer have a higher energy than those in the bulk because there are fewer bonds to neighboring molecules and therefore work must have been done on them to move them to the surface. The mea- sure of this extra energy is the surface energy per unit area ␥, given in J/m2, which depends on the particular fluids involved at the boundary (e.g., water and air). For pure FIGURE 9.14 A film of liquid is water in air the surface energy density is unusually high, ␥ ϭ 0.073 J/m2. stretched by moving the crossbar a ⌬ Associated with the increase in energy in the surface layer of fluid is a surface distance x. We use this to calcu- late the surface tension. tension. Consider the device shown in Figure 9.14 on which a liquid film, such as a soap film, is formed in air. In order to increase the surface area by sliding the cross- bar a distance ⌬x, increasing the surface area by (w⌬x), a force F is needed. The work done by this force will equal the extra surface energy, therefore we find

F¢x ϭ 2g (w¢x), (9.8) where the factor 2 enters because there are two surfaces of the fluid exposed to air. From this we find another expression for ␥,

F g ϭ , (9.9) 2w so that ␥, already seen to be the surface energy density, is also a force per unit length, with units of N/m, and is also known as the surface tension. A force per unit length is appropriate for a fluid, rather than a force per unit area, or stress, used for a solid, because the fluid surface layer is imagined to be infinitesimally thin. The surface of the liquid is sometimes said to behave like a skin or rubber sheet. This is because the surface can support small insects such as water striders skimming the surface of a pond (see Figure 9.15 and Problem 19). However, unlike a rubber sheet, when a fluid

FIGURE 9.15 A water strider glides over the water using surface tension to support itself. On the right, a dye was added to float on the water surface and when illuminated from below reveals the hydrodynamics of the strider’s motion. Note that the strider is light-seeking.

S URFACE T ENSION AND C APILLARITY 241 FIGURE 9.16 Micelle (left), vesicle (center), and planar bilayer (right) all composed of lipids.

surface is extended, additional molecules are added to the surface from the bulk fluid, and so the analogy is of limited use. If a drop of liquid is formed in air, as from a dripping faucet, the intermolecular interactions will tend to minimize the surface energy by minimizing the surface area. Because a sphere has the minimum surface area for a given volume, in the absence of other forces liquid drops are spherical. For small drops the surface tension is much larger than gravitational forces and the drops are indeed spherical. Under “weight- less” conditions, such as in a space shuttle flight, even large liquid drops are observed to be spherical. In the presence of gravity larger drops tend to get elongated vertically. We use this liquid drop idea to model the nuclei of atoms in Chapter 26 to understand the process of nuclear fission. An important related example is the formation of micelles or vesicles of lipids in water. Recall that lipids have hydrocarbon tails that are hydrophobic and polar head groups. When mixed in water at low concentrations, lipids tend to form micelles, or spherical balls with the polar groups facing water on the outside and the hydrocarbon tails buried inside (Figure 9.16). At higher lipid concentra- tions, the lipids form vesicles or spherical lipid bilayers with water both inside and outside, as shown in the center figure. These are similar to cell membranes, although cell membranes also have many associated proteins bound to the lipids. Surface tension is an important factor in the overall structure of both vesicles and micelles. In our bodies, the largest surface area in contact with air is the internal surface of the lungs. The total surface area in the lungs of an adult is tremendous, roughly 100 m2, or the size of a large room. This large surface area is possible because of a branched network of small sacs or alveoli. Figure 9.17 shows an idealized sec- tion of an alveolus taken to be spherical. The air pressure inside, Pi, is normally greater than the pressure in the pleural cavity outside, Po, and this net pressure dif- ference is balanced by the surface tension in the wall of the alveolus which we treat as an idealized elastic membrane, like a small balloon. We can relate the sur- face tension in the alveolus to the pressure difference by imagining that we divide the alveolus into two hemispheres and balance the forces acting on each separate tension hemisphere (see Figure 9.17). The net tension force pulling to the left on the right hemisphere in the figure along the circular edge of the alveolus membrane is ␲ ␥ net Fx 2 r . This force must be balanced by the net pressure force directed toward the x Ϫ right, which can be shown to equal the pressure difference Pi Po multiplied by the projected area ␲r2 (see boxed calculation). The balance of forces 2␲r␥ ϭ Ϫ ␲ 2 (Pi Po) r then implies tension FIGURE 9.17 Right hemisphere is g P Ϫ P ϭ ¢P ϭ 2 , (9.10) in equilibrium under the tension i o r forces from the left hemisphere and the pressure difference (radial forces) resulting in a net pressure which is known as Laplace’s law for a spherical membrane. This relation also holds force along the x-axis. for a spherical drop of liquid.

242 V ISCOUS F LUIDS In the lungs, both the radius of the alveoli and the pressure differ- To find the net force on the right hemi- ence vary during breathing, in part due to motion of the diaphragm (with ⌬ 2 sphere due to the pressure difference P in an area of 0.05 m ). If an alveolus were to collapse to a diameter of Figure 9.17, we need to add up the contri- about 0.2 mm, from Laplace’s law using ␥ for a water/air interface, the ⌬ ϭ butions from the normal force at each por- pressure difference would predict that a force ( PAdiaphragm # ϫ Ϫ3 # tion of the hemisphere. By symmetry, it (2 0.073/0.1 10 ) 0.05) of about 70 N would be required to should be clear that the direction of the net breathe. This is more than the weight of a newborn and is an impossibly force will be to the right because for every large force for the diaphragm to exert. To solve this problem we have a area ⌬A in the right hemisphere with nor- surfactant, a lipidÐprotein complex, present in the lungs. The addition of mal force vertical component Fy or Fz, small quantities of impurities can dramatically reduce the surface ten- there will be a symmetrically located area sion at a surface. In this case surfactants reduce the surface tension by Ϫ Ϫ with a component of Fy or Fz. Using about a factor of 15, thus greatly reducing the needed force. Premature spherical coordinates, the x-component of infants with hyaline membrane disease do not manufacture this surfac- ⌬ ϭ force due to the pressure at A is Fx tant and are prone to developing collapsed lungs. One treatment of this P cos ␪ ⌬A, where ⌬A can be written as disease involves spraying a surfactant into the lungs to temporarily sup- (r sin ␪ d␾)(rd␪). Integrating to find the port breathing. total force in the x-direction gives Suppose that a drop of liquid is placed on a plane substrate surface. Molecules on the surface of the drop have two competing forces, those p p of cohesion tending to keep the drop spherical, and those of adhesion to ϭ ¢ 2 Fx Pr df sin u cos u du the substrate surface that will tend to spread the liquid on the substrate. L0 L0 The nature of the two materials involved will determine the contact ϭ ¢Ppr2, angle u, shown in Figure 9.18. Liquids with contact angles between 0 and 90¡ are said to wet the substrate surface. Pure water wets ultraclean which was used to find Equation (9.10). glass at ␪ L 0 so that the drop spreads freely on the glass, whereas on Note that pr2 is the projected area along typical glass ␪ L 30¡. For angles larger than 90¡, for example, mercury the x-axis. on glass where the mercury beads up, the liquid does not wet the sub- strate at all. Wetting characteristics are important in our lives; we use water repellents so that water beads up and will not wet surfaces, and we add wetting agents, generally molecules with hydrophobic and hydrophilic portions, to promote better contact of a liquid with a solid surface. In biology, a most important consequence of wetting is capillary action, the rise of liquids that wet the surface of a capillary. Figure 9.19 shows a glass capillary immersed in a container of water, in which the water rises and has its characteristic meniscus and a similar tube immersed in a container of mercury showing the situa- tion for a nonwetting liquid. We can calculate the height rise h of the water in the cap- illary with radius r, by considering the surface tension that supports the weight of the water column. Because the water wets the glass at a of ␪, the vertical component of the surface tension is (see Figure 9.19 right)

F ϭ 2prg cos u, (9.11)

␲ ␪ FIGURE 9.18 The definition of the where the factor 2 r is the contact perimeter and cos accounts for the vertical contact angle for a drop on a sur- component. face; the top drop wets the surface With the weight of the water column given by ␳␲r2hg, equating these two forces and the bottom drop beads up on yields a column height of the surface, not wetting it.

2g cos u h ϭ . (9.12) θ r gr

Equation (9.12) predicts that the smaller the radius of the tube, the higher the column of fluid can rise by capillary action. It also predicts the behavior of mercury in a glass capillary because cos ␪ is negative and not only will the meniscus be inverted, but surprisingly, the column of mercury will fall below θ its level in the large container, as shown in the figure. Clearly water transport in plants and trees (sap is mostly water) is an important application of capillary action, although in this case the upper end of

S URFACE T ENSION AND C APILLARITY 243 θ the vascular system is not open to the atmosphere. Typical pore radii F in the xylem of trees is 20 ␮m, so that the maximum height rise of water in such a capillary should be about 75 cm, using a contact angle of 0¡, based on Equation (9.12). But how does water rise higher in trees, some of which are over 100 m tall? In the leaves of trees the interstitial pathways for water flow are believed to be on the order of FIGURE 9.19 (left) Capillaries 5 nm. As long as water is able to reach the leaves, it will be supported immersed in water showing the by the capillary action in the leaves, because with 5 nm pores Equation (9.12) yields meniscus and the fact that water a height of nearly 3 km, much taller than any tree. It is believed that as a tree grows, rises higher in a narrower tube; (center) capillary immersed in mer- as long as the column of water is maintained, the capillary action in the leaves is suf- cury showing the inverted menis- ficient to support the column of water. The flow of water is then regulated mainly by cus and the lower level in the evaporation from the leaves, known as transpiration, effectively producing a “negative capillary than in the surrounding pressure” that pulls water up from the soil. We know that even a vacuum cannot pull container; (right) detail showing water up to a height greater than 30 m; hence the term negative pressure, which is able surface tension force calculated in Equation (9.11). to pull water to greater heights based on capillary action. If a tree has a portion of its xylem damaged so that the water column is interrupted, then beyond a height of 75 cm there is no mechanism to restore the flow of water.

CHAPTER SUMMARY human circulatory system basically functions as two Viscosity ␩ can be defined for a Newtonian fluid as the coupled pumps that send blood to the lungs for - proportionality constant between the stress (F/A) and oxygenation and release of carbon dioxide, and to the the rate of strain (⌬v/⌬y) (where the geometry is that of capillary beds for distribution of oxygen and nutrients Figure 9.1): and collection of cellular waste products. The surface tension ␥ at the boundary surface between F ¢v two fluids (a liquid and air, e.g.) is given by the excess sur- ϭ h , (9.1) face energy per unit surface area, or equivalently by the ¢ A y force per unit length (in the geometry of Figure 9.14),

For a Newtonian fluid flowing in a cylindrical tube, F the flow rate Q is given by Poiseuille’s law, g ϭ . (9.9) 2w pPr4 Q ϭ , (9.2) The pressure difference across a spherical membrane 8hL or drop of fluid of radius r is given by Laplace’s law,

where P is the pressure difference across the tube, and g P Ϫ P ϭ ¢P ϭ 2 . (9.10) r and L are the tube radius and length, respectively. i o r The viscosity of a suspension of spherical particles ␩ ␩ s, increases from the solvent viscosity, o, as the vol- ume fraction ⌽ increases according to Capillary action causes a column of fluid of den- sity ␳ to rise a distance h ϭ ϩ £ hs ho (1 2.5 ). (9.7) 2g cos u h ϭ , (9.12) rgr Blood is a complex fluid that exhibits non- Newtonian flow and has rheological properties that are very dependent on the hematocrit, the percent of blood where ␥ and ␪ are the surface tension and contact angle volume occupied by cells (mostly red cells). The that the fluid wets the capillary surface.

244 V ISCOUS F LUIDS QUESTIONS 19. Describe in words the source of the “negative pres- 1. Give some examples of fluids with appreciable viscos- sure” that allows water to rise so high in plants and ity and try to put them in order of increasing viscosity. trees. 2. Give an argument as to why the viscosity of normal fluids should generally decrease with increasing temperature. MULTIPLE CHOICE QUESTIONS 3. Give some examples of laminar and turbulent flow of 1. The SI units for viscosity are (a) kg/(m-s), (b) kg-m2/s, fluids. (c) kg-m/s, (d) kg-s/m. 4. Assuming Poiseuille’s law applies, what would be the 2. Which has the greatest effect on the flow of fluid change in volume flow rate through a tube when the through a pipe? That is, if you made a 10% change in radius is halved? When the length is quadrupled? each of the quantities below, which would cause the When the viscosity of the liquid is doubled? When greatest change in the flow rate? (a) the fluid viscos- the pressure head is doubled? ity, (b) the pressure difference, (c) the radius of the 5. Explain how a simple timing measurement can deter- pipe, (d) the length of the pipe. mine the viscosity of a liquid in a capillary viscome- 3. In the flow of water through a capillary tube, if the diam- ter. What complications can you imagine would arise eter of the tube is tripled with no other changes, the flow in the measurement of the viscosity of suspension of rate will (a) increase by a factor of 9, (b) increase by a long polymers in a high-shear capillary tube? factor of 27, (c) increase by a factor of 16, (d) increase 6. Check that the Reynolds number is dimensionless. by a factor of 81. 7. Cigarette smokers generally have higher hematocrits 4. Which of the following is a false statement about the than nonsmokers. This is probably due to the decreased flow of a liquid in a thin vertical tube? oxygen efficiency of the red blood cells from the (a) The velocity is fastest at the center of the tube, (b) if inhaled carbon monoxide in cigarette smoke (about the tube radius is doubled the flow rate will increase by 250 cm3 per pack). What is the effect of the higher a factor of 16, (c) the ratio of the flow times for two hematocrit on the velocity of blood flow? liquids depends only on the ratio of their viscosities, 8. Explain why an aneurysm in an artery leads to a (d) the presence of suspended particles in the liquid locally elevated blood pressure. decreases the flow rate. 9. How can a plaque deposit on an artery or arteriole 5. For a given solution of particles in a solvent, the wall lead to a decreased local blood pressure and the characteristic velocity at which there is a transition collapse of that vessel? from laminar to turbulent flow is (a) proportional to 10. Hold your hands at your sides and observe a swollen the size of the particles, (b) is proportional to the vein in your arm or hand. Then raise your arm over density of the fluid, (c) is proportional to the viscos- your head. The vein will “disappear” as it shrinks in ity of the fluid, (d) is independent of the size of the diameter. Why? particles. 11. Describe, in words, the path of blood flow throughout 6. The flow of blood through a capillary requires a the human circulatory system. higher pressure where the blood enters the capillary 12. What is the difference between pulsatile and bolus than where it leaves. That is most directly related to flow of blood? (a) F/A ϭ ␩⌬v/⌬y, (b) ⌬P ϭ ␳g⌬y, (c) ⌬(P ϩ ␳gy ϩ 13. Why are artificial heart valves designed to be pas- ␳v2/2) ϭ 0, (d) ⌬(␳Av) ϭ 0. sively rather than actively controlled? 7. An object is dropped from rest at t ϭ 0 into a viscous 14. Insects that walk on water secrete antiwetting liquids fluid. Which of the following best describes the that coat their legs. How does this help them? object’s speed as a function of time? 15. Why don’t the lungs consist of two large sacs rather than huge numbers of small alveoli? Examine Laplace’s law for the answer. 16. Discuss how the competition between cohesion and adhesion determines the wetting of a material by a v v liquid. Adhesive tape (including Post-it type paper) uses this idea as well as large numbers of tiny bubbles (a) tt(b) that create vacuum suction attachments. 17. Explain the function of surfactants in our lungs. v 18. What factors control how high a fluid will rise in a v narrow capillary tube? Which ones depend on the t t fluid, the tube material, or the geometry alone? (c) (d)

Q UESTIONS/PROBLEMS 245 8. The incremental viscosity of a dilute solution of iden- 4. In giving an intravenous (IV) of saline solution to a tical particles over that of the solvent depends on all patient, the storage bag is placed 1 m above the patient’s but which of the following? (a) Particle size, (b) par- arm and attached to a hypodermic needle. If the flow ticle concentration, (c) particle shape, (d) the solvent rate out of the needle just before it is inserted into the viscosity. patient’s arm is 50 cm3/min, when it is inserted into a 9. Heart valves close in response to (a) sets of muscles, vein with a blood pressure of 20 torr (gauge pressure), (b) hydrodynamic forces, (c) vasoconstriction, how long will it take to give 1 L of saline? Assume the (d) surface tension. saline has the same physical properties as pure water. 10. Blood is called a complex fluid because (a) it has 5. Find the Reynolds number for blood pumped into the many different components, (b) it has a high viscos- 1 inch diameter aorta from the heart, using this dis- ity, (c) its viscosity depends on shear rate, (d) blood tance as the characteristic length involved. Take the plasma is a Newtonian fluid. volume of blood pumped each of 72 times per minute 11. The fundamental reason that red blood cells can flow to be 70 cm3. Is the blood flow turbulent? In fact, through small diameter capillaries at high concentra- because of the pulsatile nature of the heart pump, tions whereas plastic spheres of the same size form a blood flows into the aorta as a bolus or plug with rel- stiff “solid” is (a) the unique disk shape of the red atively little turbulence. blood cell, (b) the bolus flow of the red cells, (c) the 6. An intravenous blood plasma drip enters a vein in the flexibility of the red cell, (d) the vasoconstriction of patient’s arm from a bag raised a height h above the the capillaries. vein. If the diameter of the 5 cm long needle is 0.5 mm, 12. Heart sounds heard in a stethoscope are due to (a) tur- find the height h that results in a 5 cm3/min flow rate. bulent flow between heart chambers, (b) pulsatile (Assume the blood pressure in the arm is 18 torr.) flow in the aorta, (c) laminar flow through the heart 7. A salt solution of specific gravity 1.018 has its efflux chambers, (d) the AV shunt. time in a capillary viscometer measured to be 122.5 s 13. The shape of a droplet of liquid on a surface is due to compared to a time for distilled water of 116.4 s. a combination of (a) pressure and cohesion, (b) adhe- What is the viscosity of the salt solution? sion and cohesion, (c) capillary action and adhesion, 8. The viscosity of blood plasma is to be measured in (d) capillary action and pressure. a capillary viscometer at 37¡C. Using water as a 14. In an open tube, water can only be suctioned to rise standard, the efflux time is found to be 95 s. Predict about 10 m. In a 20 mm radius tube water will only rise the efflux time measured with blood plasma. Use about 75 cm by capillary action. How can water rise to Tables 8.1 and 9.1 and assume that the ratio of the the top of trees, sometimes over 100 m tall? (a) By densities of the two fluids is temperature-indepen- cohesive forces, (b) by adhesive forces, (c) by Laplace’s dent. Suppose this viscometer were used to try to law, (d) by transpiration generating negative pressure. measure the viscosity of whole blood. Knowing that 15. Teflon does not wet with water at all. The contact the shear forces are fairly high, would your result angle for water on teflon is (a) 0¡, (b) 90¡, (c) 180¡, be higher or lower than the low-shear value? (d) 270¡. 9. Plastic microspheres with a 5 ␮m diameter are added to water to make up a suspension. If there are 109 such spheres in a 1 cm3 volume of water, what is the PROBLEMS expected viscosity of the suspension? If the same 1. Assuming that the cream in a chocolate cream sand- numbers of 1 ␮m diameter microspheres are used in wich cookie behaves as a Newtonian fluid of 10 Pa-s the same volume of water, find the expected viscosity viscosity (probably not a great assumption), find the of this suspension. force needed to slide one of the chocolate wafers off 10. Viscosity standard solutions are to be made up from the cream at a speed of 2 mm/s if it is a 5 cm diame- distilled water and 10 ␮m diameter plastic micros- ter disk and the cream filling is 2 mm thick. pheres. If solutions of 1.05 cP, 1.1 cP, and 1.4 cP are 2. Suppose that there is a partial blockage of the aorta, desired, starting from a large volume of stock solu- which normally pumps about 5 L of blood per tion of 109 spheres per cm3, give a recipe to make up minute. If the diameter of the aorta is reduced by 100 ml of each of the three desired solutions. 30%, find the average flow rate through the diseased 11. A long fine glass capillary pipette with an inner diam- aorta. What increase in blood pressure would be eter of 0.1 mm is immersed in distilled water. How needed to obtain the normal flow rate? (Assume that high will the water rise if the glass is extremely Poiseuille’s law applies.) clean? Repeat if the contact angle is 30¡. Note that it 3. What pressure is needed to deliver saline solution is well known that the meniscus is taller for glass through a hypodermic needle with 0.3 mm inner diam- when it is extremely clean. eter and 2 cm length at a rate of 10 cm3/min. Assume 12. A spherical balloon is filled with air to a radius of the saline has the same physical properties as pure 10 cm. Find its surface tension assuming the pressure water. Also express the pressure in units of cm of H2O. inside is 5 kPa.

246 V ISCOUS F LUIDS 13. Imagine two bubbles of air of different sizes attached pressure is uniform throughout and the surface ten- via a tube with a valve as shown, all immersed in sion, really the elastic tension of the balloon material, water. With the valve closed, which bubble is at the varies with the stretch of the surface. Rank order the higher pressure? Show that unless both bubbles start surface tension, from high to low, for the labeled with the same size, when the valve is opened surface points on the balloon. tension will cause the smaller bubble to shrink and the larger one therefore to grow. This illustrates a potential problem in our lungs in inflating the alveoli, 4 or sacs connected via bronchioles. If the alveoli were not all expanded at the same rate, in theory only the 3 2 largest would form. The result of this would be a min- 1 imizing of the total surface area. Fortunately the fluid that coats our alveoli contains a surfactant that both 17. Using the previous problem, we can understand how tends to reduce the surface tension, making it easier thin-walled capillaries (thin to allow the exchange of to expand alveoli, as mentioned in the text, and also gases) can withstand the blood pressure within them. reducing the dependence of pressure on the radius so Find the elastic tension (in N/m) in a 6 ␮m radius that not all alveoli need be the same size to inflate. capillary with a blood pressure of 30 mm Hg. 18. Healthy young human arteries have a maximum elastic tension of about 500 N/m, a value that increases by more than a factor of two with age. Find the maximum pressure that such an artery can withstand before devel- oping an aneurysm, or bulge often leading to rupture, and compute how many times greater this pressure is above the normal maximum systolic pressure of 120 mm Hg. An aneurysm, or bulging of an arterial 14. Suppose an alveolus of the lung, with a radius of wall, cannot occur in a heavy artery, but only in one with 0.15 mm, is coated with pure water and devoid of sur- a weakened wall due to a connective tissue disorder. factant. What pressure would be needed to keep the alve- 19. Water striders are able to walk on the surface of water. olus inflated? Treat the alveolus as a spherical air bubble. This problem shows how they do it. Suppose that the Note that the maximum pressure that can be reached insect’s legs are nonwettable, so that the contact angle normally is about 18 mm when inhaling maximally. with water is 110¡ (assume this is from the vertical), 15. Suppose that in the previous problem the alveolus has and that the portion in contact with the water is cylin- drical. If the insect has a mass of 0.01 g (and, remem- surfactant present, reducing the surface tension to ␥ ϭ⌬ 0.03 N/m. How small can the alveolus collapse to, ber, 6 legs) use Laplace’s law for cylinders, Pr, assuming it remains spherical, and still be inflatable to find the length of each leg that must be immersed in by a strong inhalation with a maximal pressure of water to support the weight of the insect by the verti- 18 mm Hg? cal component of the surface tension. 16. Laplace’s law for cylindrical geometry is ␥ ϭ⌬Pr, where r is the radius of the cylinder; note that this is γ a factor of 2 different from that for a sphere. Consider a cylindrical balloon that is partially inflated as 110 shown. Because the balloon is in equilibrium, the

Q UESTIONS/PROBLEMS 247