You are on page 1of 34

CHAPTER

1
Hydrostatics
This chapter introduces some of the fundamental quantities involved in hydraulics, such as pressure, weight, force, mass density and relative density. It then considers the variation of pressure intensity with depth below the surface of a static liquid, and shows how the force on a submerged surface or body can be calculated. The principles outlined are used to calculate the hydrostatic forces on dams and lock gates, for example. These same principles are applied in Chapter 2 in connection with pressure measurement using piezometers and manometers, and in Chapter 3 to the analysis of oating bodies. Thus the sort of questions that are answered in this chapter are: What is meant by pressure? What is the difference between force and weight? What is the difference between mass and weight? How and why does pressure intensity vary with depth in a liquid? How can we calculate the pressure intensity at any depth? How can we calculate the force on a at immersed surface, such as the face of a dam? How can the hydrostatic force be calculated when the immersed surface is curved? Does hydrostatic pressure act equally in all directions, and if it does why? How can the buoyancy force on a body be calculated? What do we do if the liquid is stratied with layers of different density?

Understanding Hydraulics

1.1 Fundamentals 1.1.1 Understanding pressure and force

Have you ever asked yourself why a trainer will not damage a soft wooden oor, but a stiletto heel will?
The answer is because the average pressure, PAV, exerted on the oor is determined by the weight of the person, W, and the area of contact, A, between the sole of the shoe and the oor. Thus: PAV = W A
(1.1)

So, because a trainer has a at sole with a large area of contact, it exerts a relatively small pressure on the oor (Fig. 1.1). On the other hand, the sharp point of a stiletto means that much of the weight is transmitted to the oor over a small area, giving a large pressure. Similarly a drawing pin (or a thumb tack in American) creates a large, penetrative pressure by concentrating a small applied force at a sharp point. I understand that, but can you now tell me what is the difference between weight and force?

The answer is basically none. Weight is simply one particular type of force, namely that resulting from gravitational attraction. So equation (1.1) can also be written as PAV = F/A, where F is the force. This can be rearranged to give: F = PAVA
(1.2)

The unit of force is the Newton (N), named after Sir Isaac Newton, so pressure has the units N/m2. A Newton is dened as the force required to give a mass of 1 kg an acceleration of 1 m/s2. Hence:

Figure 1.1 Illustration of the pressure exerted on a oor by two types of shoe. The stiletto is the more damaging because the weight is distributed over a small area, so giving a relatively large pressure

Hydrostatics
Force = mass acceleration F = Ma

(1.3)

where M represents mass and a is the acceleration. For weight, W, which is the force caused by the acceleration due to gravity, g, this becomes: Weight = mass gravity W = Mg On Earth, gravity, g, is usually taken as 9.81 m/s2.
(1.4)

1.1.2 Understanding the difference between mass and weight


OK, so what is the essential difference between mass and weight, and why is it important?

It is important to have a clear understanding of the difference between mass and weight, because without it you will make mistakes in your calculations. The essential difference is that mass represents the amount of matter in a body, which is constant, so mass stays the same everywhere in the universe, while weight varies according to the local value of gravity since W = Mg (equation (1.4) and Fig. 1.2). So what is mass density and weight density What is meant by relative density? And how heavy is water?

(1.5)

Density, r, is the relationship between the mass, M, of a substance and its volume, V. Thus: r=M V

Figure 1.2 The concept of weight, which varies according to the local value of gravity

Understanding Hydraulics

Box 1.1

Remember
It is important to realise that water is heavy! Each cubic metre of water weighs 9.81 103 N, that is one tonne. Thus every cubic metre weighs about the same as a large car.

Figure 1.3 Illustration of the weight of water

The density of fresh water (r) is 1000 kg/m3. This can be thought of as the mass density of the water, since it gives the mass per unit volume. Alternatively, the weight (W) per unit volume may be quoted, which is the weight density, w (also called the specic weight). Using equations (1.4) and (1.5), weight density can be expressed in several ways: w =W V or w = Mg V or w = rg
(1.6)

Thus the weight density of fresh water is 1000 9.81 N/m3. Another term you may come across is the relative density (or specic gravity) of a liquid, s. This is the ratio of the density of a substance, rS, to the density of fresh water, r. Of course, the same value can be obtained by using the ratio of the weight densities (equation (1.6)), since g is the same for both substances. Thus: s = rS r or s = wS w
(1.7)

where wS is the weight density of the substance. Since s represents a ratio of the mass or weight of equal volumes of the two substances, it has no dimensions. For example, water has a relative density of 1.0 while mercury has a relative density of 13.6.

Box 1.2

Using relative density


It is important to remember that s usually has to be multiplied by the density of water before it can be used in your calculations, otherwise the answer you obtain will be wrong, both numerically and dimensionally. For example, the density of mercury (rM) is 13.6 1000 kg/m3. Quoting the relative density as 13.6 is just a shorter and more convenient way of writing this.

Hydrostatics

1.1.3 An application of what you have learned so far the hydraulic jack

You may not realise it, but you now have a sufcient understanding of hydrostatics to understand how a hydraulic jack works.

The hydraulic jack uses two cylinders (Fig. 1.4), one with a large cross-sectional area (CSA), A, and one with a small area, a. By using a handle, or something similar, a small force, f, is applied to the piston in the small cylinder. From equation (1.2), it can be seen that this generates a pressure in the liquid of PAV = f /a. Now one of the properties of a liquid is that it transmits pressure equally in all directions (more of this later), so this means that the same pressure PAV acts over the whole cross-sectional area (A) of the large piston. As a result, the force exerted on the large piston is F = PAVA (equation (1.2)). Because A > a, the output force F > f, even though the pressure of the liquid is the same. Thus the jack acts as a kind of hydraulic amplier. This simple but extremely useful effect can be used to lift weights of many tonnes while applying only a relatively small force to the input end of the jack.

1.2 Hydrostatic pressure and force

Now let us try to determine how we can work out the hydrostatic force, F, on a dam, or on a lock gate, or on the ap gate at the end of a sewer.
The term hydrostatic means, of course, that the liquid is not moving. Consequently there are no viscous or frictional resistance forces to worry about (see section 4.1). Also, in a stationary liquid there can be no shear forces, since this would imply movement. The water pressure must act at right angles to all surfaces with which the liquid comes into contact. If the pressure acted at any other angle to the surface, then there would

Figure 1.4 A hydraulic jack. The hydraulic pressure that results from applying a small force to the small piston is transmitted to the large piston, so enabling a relatively heavy load to be raised

Understanding Hydraulics

Figure 1.5 Typical examples of situations where the hydrostatic force may have to be calculated

be a component of force along it which would cause the liquid to move. However, this component is zero when the pressure is normal to the surface since cos 90 = 0. Hence in a static liquid the pressure acts at right angles to any surface. This fact comes in useful later. OK, so the pressure acts at 90 to the surface. Please can you now explain why a submarine can only dive to a certain depth, as in all those old war movies?

The answer is quite simple. The pressure intensity increases with depth. Beyond a certain depth the water pressure would crush the hull of the submarine. But what causes the pressure, and how can you calculate what it is? After all, if you were in the submarine you would want to know, right?

The weight of the water above the submarine causes the pressure. Remember, every cubic metre of fresh water equals 1 tonne, which is 9810 N (that is rg N with r = 1000 kg/m3 and g = 9.81 m/s2). This makes it quite easy to calculate the pressure. Try thinking of it like this. Imagine a large body of fresh water. Then consider a column of the liquid with a plan area of 1 m2 extending from the surface all the way to the bottom, as in Fig. 1.6. Now, suppose we draw horizontal lines at one metre intervals from the surface, so that the column is effectively separated into cubes with a volume of 1 m3. Every cube weighs 9.81 103 N. Since the pressure on the base of each of the cubes is equal to the weight of all the cubes above it divided by 1 m2 (PAV = W/A), it can be seen that the pressure increases uniformly with depth. Similarly, if the column of liquid has a total depth, d, then the total weight of all the cubes is 9.81 103 d N. Dividing this by 1 m2 to obtain the pressure on the base of the column gives 9.81 103 d N/m2. Therefore, at any depth, h, below the water surface the pressure is: P = r gh N m 2
(1.8)

Equation (1.8) shows that there is a linear relationship between pressure, or pressure intensity, and depth. This pressuredepth relationship can be drawn graphically to obtain

Hydrostatics

Figure 1.6 Variation of pressure with depth

a pressure intensity diagram like that in Fig. 1.7. This diagram shows the pressure intensity on a vertical surface that is immersed in a static liquid and which has the same height, h, as the depth of water. The arrows can be thought of as vectors: they are drawn at 90 to the surface indicating the direction in which the pressure acts, while the length of the arrow indicates the relative magnitude of the pressure intensity. When analysing a problem, a pressure intensity diagram is used to help visualise what is happening, while equation (1.8) provides the means to calculate the pressure intensity. The relationship described by equation (1.8) is very useful; it can be used to calculate the pressure at any known depth, or alternatively, to calculate the depth from a known pressure. The fact there is a precise relationship between pressure and depth forms the basis of many instruments that can be used to measure pressure, such as manometers, which are described in Chapter 2. Now one important point. Figure 1.7 only shows the pressure caused by the weight of the water. This is called the gauge pressure, and is

Figure 1.7 A pressure intensity diagram corresponding to Fig. 1.6

Understanding Hydraulics

Box 1.3

Visualising the size of units


You can easily visualise a metre, because it is just over three feet in length, and, of course, you know how long a second is. You may also be aware that a kilogramme is about 2.2 lb, that is about the equivalent of a bag of sugar. But do you know how large or small a Newton is? If you use equation (1.8) to work out the pressure at a depth of 0.3 m of fresh water you get P = 1000 9.81 0.3 = 2943 N/m2. So every time you have a bath at home, parts of your body are being subjected to almost 3000 N/m2. It does not cause any discomfort, in fact you do not even notice. So you may deduce that a Newton is a relatively small unit of force. For this reason it is frequently not worthwhile quoting a value to less than a Newton (the exception being if you are dealing with very, very small values where accuracy may be affected by rounding off).

the pressure most often used by engineers. For convenience, gauge pressure measures the pressure of the water relative to atmospheric pressure, that is it takes the pressure of the air around us as zero. Now in reality, the atmosphere exerts a pressure of about 101 103 N/m2 on everything at sea level (this is equivalent to the pressure at the bottom of a column of water about 10.3 m high, that is a head of 10.3 m of water). So if we want to obtain the absolute pressure measured relative to an absolute vacuum, that is the total pressure exerted by both the water and the atmosphere, we have to add atmospheric pressure, PATM, to the gauge pressure (Fig. 1.8). Thus the absolute pressure, PABS, is: PABS = r gh + PATM N m 2
(1.9)

A good way to think of this is that you can measure the height of a table top either from the oor, which is the most convenient way, or above sea level (ordnance datum). Similarly, it is more convenient to measure temperature above the freezing point of water than above absolute zero. Consequently in this book we will always use gauge pressures (unless stated otherwise). For future reference, note that under some circumstances, such as in pipelines, a pressure less than atmospheric may occur (Fig. 1.8). This is a negative gauge pressure, rgh, but equation (1.9) is still valid. Note also that if absolute pressure is used then the gauge pressure intensity diagram shown in Fig. 1.7 will have to have PATM added to it, as shown in Fig. 1.9. Now try Self Test Question 1.1. A short guide solution is given in Appendix 2, if you need it.

SELF TEST QUESTION 1.1


Oil with a weight density, wO, of 7850 N/m3 is contained in a vertically sided, rectangular tank which is 2.0 m long and 1.0 m wide. The depth of oil in the tank is 0.6 m. (a) What is the gauge pressure on the bottom of the tank in N/m2? (b) What is the weight of the oil in the tank?

Hydrostatics

Figure 1.8 Relationship between gauge pressure and absolute pressure

Figure 1.9 Pressure intensity diagram including atmospheric pressure

(c) If the bottom of the tank is resting not at on the ground but on two pieces of timber running the width of the tank, so that each piece of timber has an area of contact with the tank of 1.0 m 0.1 m, what is the pressure on the timber?

1.3 Force on a plane (at), vertical immersed surface


How do you work out the force on something as a result of the hydrostatic pressure? Say, something like a rectangular gate at the end of a sewer or culvert?

OK, there are two thing to remember. First of all, equation (1.2) tells us that F = PAVA, so a force is a pressure multiplied by an area. However, the second thing we have to remember is that the pressure varies with depth. So, on a vertical surface such as the gate in Fig. 1.10, the pressure at the top of the gate is rgh1. At the bottom of the gate the pressure is rgh2. Hence the average pressure on the gate is PAV = (rgh1 + rgh2)/2. Now if we multiply this by the area of the gate in contact with the water, A, we get the force, F: F = r g [(h1 + h2 ) 2] A
(1.10)

For a rectangle, (h1 + h2)/2 is the depth to the centre of the area, that is the vertical depth to the centroid, G, of the immersed surface. This depth is represented by hG, so the expression for the resultant hydrostatic force, F, becomes: F = rghGA
(1.11)

This equation can be applied to surfaces of any shape. For geometrical shapes other than a rectangle, the depth to the centroid can be found from Table 1.1. For the full derivation of equation (1.11), see Proof 1.1 in Appendix 1.

10

Understanding Hydraulics

Figure 1.10 A vertical gate at the end of a sewer which discharges to a river. The gate hangs from a hinge at the top: (a) side view, (b) front view, (c) pressure intensity diagram. Note that only the part of the pressure intensity diagram at the same depth as the gate contributes to the hydrostatic force acting on it

Table 1.1
Shape

Geometrical properties of some simple gures


Dimensions Location of the centroid, G D/2 from base D/3 from base centre of the circle 4R/3p from base Second moment of area, IG LD3/12 LD3/36 pR4/4 0.1102R4

Rectangle Triangle Circle Semicircle

breadth L height D base length L height D radius R radius R

The next paragraph can be helpful in some circumstances, since it reconciles what can appear to be different ways to solve a particular problem. However, you may omit it the rst time you read the chapter, or if it confuses you. From equation (1.10), the resultant force, F = average pressure intensity area of the immersed surface (A). For simple, at surfaces like that in Fig. 1.10, the average pressure intensity is (rgh1 + rgh2)/2. If A = DL, then equation (1.10) can be written as F = rg[(h1 + h2)/2]DL. The same expression can be obtained by calculating the area of the trapezoidal pressure intensity diagram in contact with the gate, rg[(h1 + h2)/2]D and multiplying by the length of the gate, L. This can sometimes provide a useful check that what you are doing is correct, or a means of remembering the equation. However, your best approach initially is usually to go straight to equation (1.11).

Hydrostatics

11

Box 1.4

Remember
Whenever you are faced with calculating the horizontal hydrostatic force on a plane, vertical immersed surface, the equation F = rghGA is the one to use. This simple equation can solve a lot of problems. We will also use it later on when we progress to the force on inclined and curved immersed surfaces. Remember that A is the area of the immersed surface in contact with the liquid.

1.4 Location of the resultant force on a vertical surface

How do you know where the resultant force, F, acts?


Yes, there is a way of calculating where the resultant force acts, and normally you would work this out at the same time as the magnitude of the force itself. However, the proof is a bit complicated, so I have put it in Appendix 1 (the second half of Proof 1.1). You can go through it later if you want to. For the time being, though, let us try to deduce something about where the force must act. Consider the dam in Fig. 1.11. In this case the pressure intensity diagram is triangular, since the gauge pressure varies from zero (atmospheric pressure) at the surface to rgh at the bottom. The average pressure intensity on the dam is therefore (0 + rgh)/2 or rgh/2. This pressure occurs at G, half way between the water surface and the bottom of the dam. But where would the resultant force act? At G, half way down? Above? Below? Can you deduce where it would be?

I assume that there must be some way of working it out?

Figure 1.11 Pressure intensity on a dam. G is the centroid of the wetted area, P is the centre of pressure where the resultant force acts

Think of it this way. The resultant force on the dam is the result of the average pressure intensity acting over the area of the dam face in contact with the water. The longer the arrows of the pressure intensity diagram, the greater the pressure. The larger the area of the pressure intensity diagram, the greater the force.

12

Understanding Hydraulics

Box 1.5

Note that the centre of pressure, P, is always below the centroid, G, of the surface in contact with the water. In many problems it is not obvious where P is located, so this has to be calculated using equation (1.12). However, as the depth of immersion of the surface increases, P moves closer to G. This is apparent from equation (1.12): the distance between P and G is (hP - hG). If A and IG have constant values, then the equation can be rearranged as (hP - hG) = C/hG where C represents the value of the constants. Thus (hP - hG) decreases as hG increases.

Look at the triangular area that forms the top half of the pressure intensity diagram, and compare it with the area of the trapezoidal bottom half. The area of the bottom part of the diagram is much larger, indicating that the resultant force would act below half depth. In fact, the resultant force acts horizontally through the centroid of the pressure intensity diagram. For the triangular pressure intensity diagram in Fig. 1.11, this is located at h/3 from the base (but note that this is only the case when the pressure intensity diagram is triangular). The point, P, at which the resultant force acts is called the centre of pressure (Fig. 1.11). With more complex problems, like that in Fig. 1.10, there is no simple rule to give the location of P, but if hP is the vertical depth to the centre of pressure then this can be calculated from: hP = ( IG AhG ) + hG
(1.12)

where the value in the brackets gives the vertical distance of P below the vertical depth to the centroid of the surface, hG. The appropriate expression for the second moment of area calculated about an axis through the centroid, IG, can be found from Table 1.1. For a rectangle IG = LD3/12, where L is the length of the body and D its height. A is the surface area of the body. The derivation of equation (1.12) can be found in Appendix 1. Examples 1.1 and 1.2 show how equations (1.11) and (1.12) are used to solve a couple of typical problems, one involving the ap gate at the end of a sewer and the other a lock gate. Study these carefully and then try Self Test Question 1.2 (a short solution is given in Appendix 2).

SELF TEST QUESTION 1.2


A rectangular culvert (a large pipe) 1.8 m wide by 1.0 m high discharges to a river. At the end of the culvert is a rectangular gate which seals off the culvert when the river is in ood (as in Fig. 1.10). The gate hangs vertically from hinges at the top. If the ood level in the river rises to 1.9 m above the top of the gate, calculate the magnitude and location of the resultant hydrostatic force on the gate caused by the water in the river.

EXAMPLE 1.1
A rectangular gate is 2 m wide and 3 m high. It hangs vertically with its top edge 1 m below the water surface. (a) Calculate the pressure at the bottom of the gate. (b) Calculate the

Hydrostatics

13

resultant hydrostatic force on the gate. (c) Determine the depth at which the resultant force acts. (a) From equation (1.8), P = rgh Therefore P = 1000 9.81 (3 + 1) = 39.24103 N m2 (b) From equation (1.11), F = rghGA
G

Now hG = 1+ (3 2) = 2.50 m A = 2 3 = 6m2 Thus F = 1000 9.81 2.50 6 = 147.15 103 N (c) From equation (1.12)

h P = ( I G AhG ) + hG where I G = LD 3 12 = 2 33 12 = 4.50 m4 A and hG are as above so hP = (4.50 6 2.50) + 2.50 = 2.80 m

Figure 1.12

EXAMPLE 1.2
A lock on a canal is sealed by a gate that is 3.0 m wide. The gate is perpendicular to the sides of the lock. When the lock is used there is water on one side of the gate to a depth of 3.5 m, and 2.0 m on the other side. (a) What is the hydrostatic force of the two sides of the gate? (b) At what height from the bed do the two forces act? (c) What is the magnitude of the overall resultant hydrostatic force on the gate and at what height does it act? (a) Using F = rghGA F1 = 1000 9.81 (3.5 2) (3.5 3.0) = 180.26 103 N F2 = 1000 9.81 (2.0 2) (2.0 3.0) = 58.86 103 N (b) Since both pressure intensity diagrams are triangular, both forces act at onethird depth from the bed: Y1 = 3.5 3 = 1.17 m Y2 = 2.0 3 = 0.67 m (c) Overall resultant force FR = F1 - F2 FR = 121.40 103 N Taking moments about O to nd the height, YR, of the resultant: 121.40 103 YR = 180.26 103 1.17 - 58.86 103 0.67 YR = 1.41 m above the bed.

Figure 1.13

14

Understanding Hydraulics
The value of YR obtained in part (c) of the above example may have surprised you. Possibly you expected YR to be somewhere between 0.67 m and 1.17 m, whereas it is actually 1.41 m. This is a situation where the pressure intensity diagrams (which are not really needed to conduct the calculations) can be used to visualise what is happening. In Fig. 1.13 the slope of the two pressure intensity triangles is the same, since the water has the same density on both sides of the gate. Thus if the triangle on the right is subtracted from the triangle on the left, the Figure 1.14 Net pressure intensity result is as in Fig. 1.14. This is the net presdiagram for Example 1.2 sure intensity on the gate. The diagram is more rectangular than either of the triangles so, employing a similar argument to that used with Fig. 1.11, this indicates that YR would be higher above the base than either Y1 or Y2.

Figure 1.15 The dam on the bottom left of the photograph is holding back a considerable quantity of water. The force exerted by the water on the structure must be calculated before the dam can be designed. Many lay people believe, incorrectly, that the greater the volume of water stored behind the dam, the larger the force on the structure. This is not the case. Equation (1.8) indicates that the pressure on the dam is related to the depth of water, while the force is the product of the average pressure and the area of the dam in contact with the water (equation (1.2))

Hydrostatics

15

1.5 Force on a plane, inclined immersed surface


I understand how to work out the force on a at vertical surface, but how about one that is inclined at an angle to the water surface? Surely this is much more difcult?

The answer is no. The calculations are still very simple and almost identical to those above. There are three things that you should remember when analysing these situations: (1) The resultant force acts at right angles to the immersed surface. (2) The hydrostatic pressure on the inclined surface is still caused only by the weight of water above it, so P = rgh. (3) When calculating the location of the resultant force on an inclined surface, always use equation (1.13) (never equation (1.12), see below). To illustrate simply that the resultant force can be calculated in the same way as for a vertical surface, consider this. The pressure at the top of the rectangular, inclined surface in Fig. 1.16a is rgh1 while that at the bottom is rgh2. Thus the average pressure intensity on the surface is rg(h1 + h2)/2, or rghG since hG = (h1 + h2)/2. The resultant force is the average pressure intensity multiplied by the area of the surface, and since the pressure acts at right angles to the inclined surface the actual area, A, should be used. Thus F = rghGA, as in equation (1.11). Note that the inclination of the surface is automatically taken into account by the value of hG. For example, if h1 in Fig. 1.16a is xed, and the surface rotated upwards about its top edge, then hG will decrease so that hG = h1 when it is horizontal. Similarly, the maximum possible value of hG would be obtained when the surface is vertical. One other important point, the resultant force on the inclined surface, F, has components in both the vertical and horizontal directions. These can be calculated separately, as in section 1.6 and Example 1.4, but the procedure outlined above is quicker for at (plane) surfaces. To calculate the location of the resultant force, the following equation should be used:

G G

Figure 1.16 (a) Force on an inclined surface. (b) When the surface is inclined always use the dimensions LG and LP with equation (1.13) (never the vertical dimensions hG and hP with equation (1.12))

16

Understanding Hydraulics

Box 1.6

Using equations (1.12) and (1.13)


Remember that when you have an inclined surface, always use equation (1.13) to nd the location of the resultant force. You can then calculate the vertical depth of the centre of pressure, P, below the surface from LP if you want to (see Example 1.3). Never try to do this by using equation (1.12) instead of equation (1.13). The reason for this is that I G is calculated in the plane of the surface. For example, with a rectangular inclined surface, I G is still taken as LD3/12 where D is the actual inclined dimension of the surface, so the remainder of the terms in equation (1.13) must have the same orientation for consistency (see the derivation of the equation in Appendix 1). The same argument applies to vertical surfaces and equation (1.12).

LP = ( IG ALG ) + LG

(1.13)

This is similar to equation (1.12), but the inclined lengths, LP and LG, are used to denote the location of the centre of pressure and centroid of surface (Fig. 1.16b), not the vertical depths.

EXAMPLE 1.3
A sewer discharges to a river. At the end of the sewer is a circular gate with a diameter (D) of 0.6 m. The gate is inclined at an angle of 45 to the water surface. The top edge of the gate is 1.0 m below the surface. Calculate (a) the resultant force on the gate caused by the water in the river, (b) the vertical depth from the water surface to the centre of pressure. (a) Vertical height of gate = 0.6 sin 45 = 0.424 m Vertical depth to G = hG = 1.000 + 0.424/2 = 1.212 m Area of gate, A = pD2/4 = p0.62/4 = 0.283 m2 F = rghGA = 1000 9.81 1.212 0.283 = 3365 N

Figure 1.17 An inclined, circular gate at the end of a sewer

Hydrostatics
(b) Slope length to G, LG = 1.212/sin 45 = 1.714 m For a circle (Table 1.1) I G = pR4/4 = p(0.3)4/4 = 0.0064 m4 LP = ( I G ALG ) + LG = (0.0064 0.283 1.714) + 1.714 = 1.727m Vertical depth to P, hP = LP sin 45 = 1.727 sin 45 = 1.221 m

17

Figure 1.18 This vertical lift gate on the Old Bedford River provides another example of where the engineer may be required to calculate the resultant hydrostatic force. If the horizontal force is large it may be difcult for a vertical lift gate to slide up and down, the gate being pushed hard against the guide channels. In the Fens of East Anglia much of the drainage is controlled by man, using pumps and sluice gates like the one above

18

Understanding Hydraulics

1.6 Force on a curved immersed surface

I suppose that you are now going to tell me that working out the force on a curved surface is just as easy as calculating the force on a at or inclined surface?
Well, the calculations are perhaps a little longer, but no more difcult. Let me clarify this by breaking the analysis of the force on an immersed curved surface down into steps. (1) The resultant force (F) acts at right angles to the curved surface. This force can be thought of as having both a horizontal (FH) and a vertical (FV) component (Fig. 1.19). (2) To calculate the horizontal component of the resultant force (FH), project the curved surface onto a vertical plane, as in Fig. 1.20. This effectively is what you would see if you looked at the curved surface from the front. Calculate the force on this projected vertical surface as you would any other vertical surface using FH = rghGA, where A is the area of the projected vertical surface (not the area of the actual curved surface). (3) Calculate the vertical component of the resultant force (Fig. 1.21) by evaluating the weight of the volume (V ) of water above the curved surface, that is: FV = rgV (4) The resultant force, F, is given by: F = ( FH + FV
2 2 1 2

(1.14)

(1.15)

(5) The direction of the resultant force (Fig. 1.22) can be found from: tan f = FV FH This gives the angle, f, of the resultant to the horizontal. Remember, the resultant also acts at 90 to the curved surface, so it passes through the centre of curvature (for example, the centre of the circle of which the surface is a part).
(1.16)

Figure 1.19 Pressure intensity on a curved surface. F passes through the centre of curvature, C

Figure 1.20 Projection of the curved surface onto a vertical plane

Hydrostatics

19

Figure 1.21 The vertical component of force, FV, caused by the weight of water above the surface

Figure 1.22 The direction of the resultant force, F, which must also pass through C

(6) The above steps enable the resultant force on the upper side of the surface to be calculated. Always remember that there is an equal and opposite force acting on the other side of the surface. This fact comes in useful later, because it is always easier to calculate the force on the upper surface, even if this is not the surface in contact with the water.

EXAMPLE 1.4
A surface consists of a quarter of a circle of radius 2.0 m (Fig. 1.22). It is located with its top edge 1.5 m below the water surface. Calculate the magnitude and direction of the resultant force on the upper surface.

Step 1

Project the curved surface onto a vertical plane and calculate FH FH = rghGA where A is the area of the projected vertical surface. Since the length of the gate is not given, calculate the force per metre length with L = 1.0 m. Thus A = 2 1.0 = 2.0 m2 per metre length The value of hG is that for the projected vertical surface: hG = 1.5 + (2.0/2) = 2.5 m. F H = 1000 9.81 2.5 2.0 = 49.05 103 N m

Step 2

Calculate FV from the weight of water above the surface FV = rgV where V is the volume of water above the curved surface. Again using a 1 m length: V = (1 4 p 2.02 1.0) + (2.0 1.5 1.0) = 6.14 m3 per metre length F V = 1000 9.81 6.14 = 60.23 103 N m.

Step 3

Calculate the magnitude and direction of the resultant force F = (FH + FV


2 2 12

= 103 (49.052 + 60.232 )

12

= 77.68 103 N m.

f = tan-1(FV FH ) = tan-1(60.23 49.05) = 50.8 . The resultant passes through the centre of curvature, C, at an angle of 50.8.

20

Understanding Hydraulics

Figure 1.23

SELF TEST QUESTION 1.3


An open tank which is 4.0 m wide at the top contains oil to a depth of 3.4 m as shown in Fig. 1.24. The bottom part of the tank has curved sides which have to be bolted on. To enable the force on the bolts to be determined, calculate the magnitude of the resultant hydrostatic force (per metre length) on the curved surfaces and its angle to the horizontal. The curved sections are a quarter of a circle of 1.5 m radius, and the oil has a relative density of 0.8.

Figure 1.24 Tank for Self Test Question 1.3.

I understand Example 1.4, but when you described the steps used to analyse the force on a curved surface, in point 6 you said something about always analysing the upper side of the surface. You said that we should do this even if the upper side of the surface was not in contact with the water. How can this be right? No water, no hydrostatic force I would have thought.

I suppose this is one of the tricks you have to learn to make hydraulics easy. Think of it like this. The curved surface in Fig. 1.25 is an imaginary one, drawn in a large body of static liquid. Now it is possible to calculate the force on the upper side of this imaginary surface

Hydrostatics

21

using the same procedure as in Example 1.4. However, the surface is only imaginary, so what resists this force? Something must because the liquid is static, that is not moving. The answer is that there is an equal and opposite force acting on the underside of the imaginary surface, so that this balances the force on the top. It does not matter which force you calculate, because they are numerically equal, but it is easier to calculate that on the upper surface. The same is true with real Figure 1.25 Equal and opposurfaces. Remember this when you encounter site forces on a surface problems like Example 1.5 with air on the upper surface and water underneath. Something to note from Example 1.5 is that the vertical component of the resultant force acts upwards, which means that it is a buoyancy force. Sometimes there is a tendency to think of a buoyancy force as being different from the hydrostatic force, but in fact they are the same thing. The buoyancy force on a body, such as a ship, is the result of the hydrostatic pressure acting on the body. This will be explored in more detail in section 1.7.

EXAMPLE 1.5
A radial gate whose face is part of a circle of radius 5.0 m holds back water as shown in Fig. 1.26. The sector of the circle represented by the gate has an angle of 30 at its centre. Water stands to a depth of 2.0 m above the top of the upstream face of the gate. The other side of the gate is open to the atmosphere. Determine the magnitude and direction of the resultant hydrostatic force. The gate is 3.5 m long.

Step 1

Project the curved surface onto a vertical plane and calculate FH Vertical height of projection = BC = 5.0 cos 60 = 2.5 m.

Figure 1.26

22

Understanding Hydraulics
hG = 2.0 + (2.5 2) = 3.25m, and A = 2.5 3.5 = 8.75m2 . FH = rghGA = 1000 9.81 3.25 8.75 = 278.97 103 N.

Step 2

Calculate the vertical component, FV, from the weight of water above the surface In this case calculate the weight of water that would be above the gate if it was not there, that is the weight of the water displaced by the gate. This is shown in the diagram as AEFH. The width of this area, DE, can be calculated as follows: AB = 5.0 sin 60 = 4.33 m, so DE = 5.00 - 4.33 = 0.67 m. The area of ADE (and subsequently AEFH) can be found using geometry, as follows. Area sector ACE = (30/360)ths of a 5.0 m radius circle = (30/360)p5.02 = 6.54 m2. Area triangle ACD = (1/2) 4.33 2.5 = 5.41 m2. Area ADE = 6.54 - 5.41 = 1.13 m2. Therefore, the total area AEFH = 1.13 + (0.67 2.00) = 2.47 m2. Volume of water displaced, V = 2.47 3.5 = 8.65 m3. FV = rgV = 1000 9.81 8.65 = 84.86 103 N.

Step 3

Calculate the magnitude and direction of the resultant force F = (FH + FV


2 2 12

= 103 (278.972 + 84.862 )

12

= 291.59 103 N.

f = tan-1(FV FH ) = tan-1(84.86 278.97) = 16.9 Resultant acts at 16.9 to the horizontal passing upwards through the centre of curvature, C.

1.7 Variation of pressure with direction and buoyancy


We have already discussed the fact that hydrostatic pressure acts at right angles to any surface immersed in it, so it follows that on the underside of a horizontal surface the resultant force is acting vertically upwards. This is a buoyancy force, and it is caused simply by the hydrostatic pressure on the surface. Try thinking it through like this.

Imagine a sphere some distance below the water surface as in Fig. 1.27. The hydrostatic pressure acts at 90 to the surface of the sphere. Looking at this two-dimensionally, as in the diagram, then the smallest pressure intensity is rgh1 at the top, and the largest is rgh2 at the bottom. Now, consider what would happen if the diameter of the sphere gradually decreased so that the difference between h1 and h2 decreased. This would cause the two pressures rgh1 and rgh2 to become closer numerically. If the diameter of the sphere continued to decrease until it

h3

Figure 1.27 Pressure on a sphere

Hydrostatics

23

became innitesimally small then the difference between h1 and h2 would be negligible so that rgh1 = rgh2. By the same argument, the pressure intensity in any other direction, such as rgh3 acting horizontally, would also have the same value (see Proof 1.2, Appendix 1). Thus the pressure at a point in a static liquid acts equally in all directions, up, down, sideways or whatever.

Thats all very interesting, but does it have any practical purpose, and how can I work out the value of the buoyancy force?
Yes it has a practical purpose, and working out the value of the buoyancy force is quite easy. In fact you can do so using what you have already learnt. Let me illustrate by using a similar situation to the sphere in Fig. 1.27, but this time we will make the body a cube because it simplies the calculations. The cube is shown in Fig. 1.28. The pressure intensities on the vertical sides cancel each other out, so only the pressure acting on the top and bottom faces need be considered. Let the area of each face of the cube be A. Then: Assuming the top and bottom faces are in a horizontal plane, then the pressure is constant over the face so: Pressure on the top face = rgh1 Pressure on the bottom face = rgh2 The force on the face is equal to the pressure multiplied by the area of the face, A. So: Force on the top face = rgh1A Force on the bottom face = rgh2A Since h2 > h1 there will be a net force acting vertically upwards, F. This is: F = rgh2 A - rgh1A = rg(h2 - h1)A Now (h2 - h1)A is the volume of the cube, V, so: F = rgV
(1.14)

Figure 1.28 Buoyancy force, F

Box 1.7

Remember
The buoyancy force, F, acts vertically upwards through the centre of gravity of the displaced liquid (such as the centre of the cube). The point at which F acts is called the centre of buoyancy, B. The force, F, is equal to the weight of the volume of liquid displaced by the body, that is rgV. This is known as Archimedes Principle. Now go back and look at Step 2 of Example 1.5. You should be able to see that a buoyancy force is just the vertical force caused by hydrostatic pressure. See also Chapter 3 and Box 3.1.

24

Understanding Hydraulics

When we analysed the buoyancy force on the cube in Fig. 1.28 we only considered the hydrostatic forces acting vertically on it. The weight of the cube was irrelevant. However, if we wanted to know whether or not the completely immersed cube would oat or sink, we would have to compare the weight of the cube (W ) with the buoyancy force (F ), remembering that weight is a force.

W = weight density of cube material volume = rS gV N acting vertically downwards. F = weight of liquid displaced by the cube = rgV N acting vertically upwards. Since g and V are the same, it follows that if the density of the substance, rS, that forms the cube is greater than the density of the liquid, r, then the cube would sink (W > F ). Conversely, if rS < r, then the cube would oat (F > W ). If rS = r then the cube has neutral buoyancy and would neither oat nor sink, but would stay at whatever depth it was located (F = W ). The analysis above explains why a concrete or steel cube would sink, and a cork or polystyrene cube would oat. However, this assumes that the cube is solid. If the cube was hollow, its average density would have to be used in the calculations, not the density of the material from which it was made. Submarines provide an interesting example, because they must be able to sink and, more importantly, rise to the surface again. This can be achieved by adjusting the average density of the submarine, by changing its weight by admitting or expelling water from tanks on the outside of the hull. Floating bodies, such as ships and the pontoon in Example 1.7, are quite easy to analyse. If the depth of immersion is constant, then obviously W and F in Fig. 1.29 are exactly equal (otherwise the body would move Figure 1.29 Floating up or down). Hence the starting point for many calculabody tions involving oating bodies is: W =F W = r gV

or

(1.17)

Thus a oating body of weight W displaces a volume of water (V ) that has a weight (rgV ) equal to its own. Since W = Mg this can also be written as: Mg = r gV or M = rV
(1.18) (1.19)

Therefore it is also true to say that a oating body of mass M displaces a volume of water (V) that has a mass (rV ) equal to its own. Of course, equation (1.19) is a rearrangement of equation (1.5). Remember to use W with the weight density (rg) and M with the mass density ( r). Typically the bodys weight or mass is known, so the relationships above allow the volume of water (V ) displaced by a oating body to be calculated. Then for pontoons which are rectangular in plan and cross-section like those in Fig. 3.1: depth of imersion = V/plan area
(1.20)

By now it should be apparent that a solid steel cube sinks, but a ship made from steel plates oats because it is hollow and can displace a much larger volume of water (V ) that has a

Hydrostatics

25

Figure 1.30 Lock gates provide another example of where it may be necessary to calculate hydrostatic forces. The buoyancy, depth of immersion and freeboard (the distance from the deck to the waterline) of the barge may also be the subject of an engineers calculations

mass (or weight) equal to that of the ship. This is why we say that a ship has a displacement of 10 000 tonnes, for instance. When W = F the depth of immersion is constant, but if W is increased by adding cargo the ship settles deeper in the water, increasing its displacement and consequently F, until W = F again. From 1876 onwards, British ships have had a Plimsoll line painted on their hull to indicate the maximum safe loading limit. Since the density of water changes according to temperature and salt content, the Plimsoll line includes marks for sea or fresh water, winter or summer, in tropical or northern waters.

26

Understanding Hydraulics

Box 1.8

Try this amaze your friends


Get an empty zzy drink bottle, ll it completely with water and put a sachet of ketchup in it (Fig. 1.31a). You need one that just oats, so you may have to try a few different types until you nd one that works. Now challenge your friends to concentrate their minds and use the power of thought to make the sachet sink. Unless they really do have telekinetic powers, they wont be able to do it of course. Now heres the trick. When it is your turn, make sure you have your hands around the bottle, and gently squeeze it. Try to disguise the fact you are doing this. If you squeeze hard enough the sachet will sink, and you can claim to have a better brain than all of your friends combined. The reason the sachet sinks is as follows. A body in water has two forces acting on it: its weight (W) acting vertically down and the buoyancy force (F = rgV ) acting vertically up. The weight of the sachet cannot change, so W is constant. However, F depends upon the volume (V ) of water displaced by the sachet. When you squeeze the bottle you are exerting pressure on the water inside. The water is incompressible, but the air in the sachet can be compressed. So by compressing the air, V is reduced and so is F. When W > F the sachet sinks. When you stop squeezing F > W so the sachet rises. Human divers can control their buoyancy and move up and down like this, either by inating or deating their dry suits or by controlling the amount of air in their lungs. Usually a Cartesian diver consists of a small length of open ended glass tubing with a bubble at one end (Fig. 1.31b). It can be used instead of the sachet and works in the same way.
W

Air

(a)

(b)

Figure 1.31 (a) Alternative Cartesian diver using a sachet of sauce. Squeezing and releasing the bottle makes the diver sink and then rise. (b) Conventional glass diver

Hydrostatics

27

EXAMPLE 1.6
A pipe which will carry natural gas is to be laid across an estuary which is open to the sea. The weight of the pipe is 2360 N per metre length and its outside diameter is 1.0 m. The weight of the gas can be ignored. The density of sea water is 1025 kg/m3. Determine whether the pipe will remain on the sea bed or oat. If it does oat, what force would be required to hold the pipe on the sea bed? The maximum buoyancy force occurs when the pipe is fully submerged. Thus: Buoyancy force, F = rgV. V = 1 p (1.0) 4 = 0.785 m3 m length. F = 1025 9.81 0.785 = 7893 N m. Weight of pipe, W = 2360 N m. Therefore, net force on pipe = (7893 - 2360) = 5533 N m The net force acts upwards since F > W. The pipe would oat and a force of at least 5533 N/m would be required to hold it down.
2

Figure 1.32

EXAMPLE 1.7
A pontoon which is being used to conduct some construction work on a pier built into the sea has a mass of 50 tonnes (1 tonne = 1000 kg). The pontoon is rectangular in plan and crosssection. Its length is 10 m, its width 5 m, and its sides are 2 m high. The density of sea water (rSW) is 1025 kg/m3. (a) Determine the volume of water displaced by the pontoon. (b) Determine the depth of immersion and the freeboard of the pontoon. (c) Determine the buoyancy force on the pontoon. (a) A oating body displaces its own mass (or weight) of water, so from equation (1.19), volume displaced = mass of pontoon/rSW = 50 000/1025 = 48.78 m3.

(b) Depth of immersion (the depth in the water) = volume displaced/plan area = 48.78/(10 5) = 0.98 m. Freeboard = (2.00 - 0.98) = 1.02 m. (c) Buoyancy force = weight of water displaced = weight of pontoon = 50 000 9.81 = 490.5 103 N. (Alternatively F = rgV = 1025 9.81 48.78 = 490.5 103 N)

Figure 1.33

28

Understanding Hydraulics

1.8 The hydrostatic equation


The hydrostatic equation is really a statement of what, by now, should be obvious to you. Nevertheless, it can be useful, so the meaning of the equation is considered below while the derivation of the equation can be found in Appendix 1. Basically, the hydrostatic equation states that the change in pressure intensity between two levels of a homogeneous (uniform) liquid is proportional to the vertical distance between them. Consider points 1 and 2 at some distance below the surface as in Fig. 1.34. This time let us measure the depth of the points from the bottom (not from the surface) and let these distances be denoted by z1 and z2. Pressure at point 1, P1 = rgh1 = rg (d - z1 ) Pressure at point 2, P2 = rgh2 = rg (d - z2 ) The difference in pressure between the two points is (P2 - P1) where:

( P2 - P1 ) = rg (d - z2 ) - pg (d - z1 ) = rg (d - z2 - d + z1 ) = rg ( - z2 + z1 ) ( P2 - P1 ) = - rg ( z2 - z1 )

(1.21)

Equation (1.21) shows that the difference in pressure between two points is equal to the vertical distance (z2 - z1) between them. However, the equation is more useful when rearranged so:

Figure 1.34 Pressure intensity at two points

( P2 r g ) + z2 = ( P1 r g ) + z1

(1.22)

This equation contains four of the six terms of the Bernoulli (or energy) equation that will be discussed in Chapter 4. The two terms that are missing from the Bernoulli equation are the velocity heads (V 2/2g), which is logical since the velocity (V ) is zero in a static liquid. When we start considering pressure measurement using manometers in the next chapter we will be using equations (1.21) and (1.22), or at least the meaning of the equations if not the actual equations themselves. Perhaps you can see from the equations that if you know the pressure (say P1) at some point in a static liquid, then you can calculate the pressure (P2) at any other point so long as you can measure the vertical distance between them. Manometers are designed to enable the pressure difference (P2 - P1) to be determined from the difference in the height of two columns of liquid (z2 - z1), knowing the weight density of the liquid rg. Do not worry if you do not fully understand this at the moment, since manometers are explained in the next chapter.

1.9 Stratied uids


How do you calculate the pressure if you have two liquids of different density? Does this make things more difcult?

Hydrostatics

29

Box 1.9

The equal level, equal pressure principle


One nal thing about the hydrostatic equation, which again is a statement of the obvious, is that at a constant depth (or height z in the case of Fig. 1.35) the pressure is constant. It has to be since P = rgh. However, this gives rise to the equal level, equal pressure principle. This simply states that if you draw a horizontal line in a continuous body of static, uniform uid then the pressure is the same anywhere on that line. The meaning and signicance of this will be clearer if you look at Fig. 1.35. Again, this principle is used with manometers and will be used in the next chapter. However, remember the liquid must have a uniform density (otherwise see section 1.9).

Figure 1.35 Equal level, equal pressure principle. The broken line is horizontal and the liquid has a constant density, so the hydrostatic pressure is constant along the line and P1 = P2 = P3

Well, again this is nothing new. If you look back to section 1.2 you will see that we discussed the fact that hydrostatic pressure is caused simply by the weight of the liquid above the point (or surface) that we are considering. This is still true when you have two or more liquids of different densities. All you have to do is work out the weights (or pressures) of the liquids one at a time then add them together. Let us analyse the situation in Fig. 1.36. Say that the column of liquid has a plan area of A m2. The weight of the upper block of liquid is given by: W1 = weight density volume = r1gh1A Similarly, the weight of the lower block is: W2 = r2gh2A Total weight WT = W1 + W2 = r1gh1A + r2gh2A Now equation (1.1) told us that: pressure = weight area

Figure 1.36 A stratied liquid with layers of density r1 and r2

30

Understanding Hydraulics

Box 1.10

Remember
Because the liquid is stratied and has two different densities, the pressure intensity diagram does not have the same gradient over the whole depth (as it did in Fig. 1.7, for instance). Instead, there is a change in gradient at the interface between the two liquids. However, within a particular liquid the gradient is uniform. Figure 1.37b in Example 1.8 provides an illustration of this.

so the total pressure, PT, at the base of the column of liquid is: PT = WT A PT = r1 gh1 + r2 gh2

(1.23)

We have analysed many different situations in this chapter. To help you to remember

how to approach the different types of problem, I have provided a summary for you at the end of the chapter. This may prove useful when revising or when tackling the revision questions.

EXAMPLE 1.8
A tank with vertical sides contains both oil and water. The oil has a depth of 1.5 m and a relative density of 0.8. It oats on top of the water, with which it does not mix. The water has a depth of 2.0 m and a relative density of 1.0. The tank is 3.0 m by 1.8 m in plan and open to the atmosphere. Calculate (a) the total weight of the contents of the tank; (b) the pressure on the base of the tank; (c) the variation of pressure intensity with depth; (d) the force on the side of the tank. (a) WT = ( r1gh1 + r2gh2)A Plan area A = 3.0 1.8 = 5.4 m2 WT = (0.8 1000 9.81 1.5 + 1.0 1000 9.81 2.0)5.4 = (11 772 + 19 620)5.4 = 169 517 N (b) Total pressure at base of tank = WT/A = 169 517/5.4 = 31 392 N/m2 (c) Pressure at the surface = atmospheric = 0 Pressure at the bottom of the oil = r1gh1 = 11 772 N/m2 Total pressure at the bottom of the tank = 31 392 N/m2 The pressure intensity diagram is shown in Fig. 1.37b. (d) The side of the tank is 3.0 m long. The force on the side of the tank can be obtained from equation (1.2) by multiplying the area of the tank in contact with each of the liquids by the average pressure intensity of the particular liquid.

Hydrostatics

31

Figure 1.37 (a) Tank containing a stratied liquid, and (b) the corresponding pressure intensity diagram

Average pressure of the oil = (0 + 11 772)/2 = 5886 N/m2 Force due to the oil = 3.0 1.5 5886 = 26 487 N Average pressure of the water = (11 772 + 31 392)/2 = 21 582 N/m2 Force due to the water = 3.0 2.0 21 582 = 129 492 N Total force on the side = 26 487 + 129 492 = 155 979 N

32

Understanding Hydraulics

Summary

G G

G G

G G G G G G G G

G
G

G G

G G G G G G G G

G G

Hydrostatics

33

Revision questions
1.1 Dene clearly what is meant by the following, and give the appropriate units in each case: (a) pressure; (b) force; (c) weight; (d) gravity; (e) mass; (f ) mass density; (g) weight density; (h) relative density; (i) hydrostatic pressure; (j) buoyancy force. 1.2 (a) Explain what is meant by gauge pressure and absolute pressure. (b) What is the approximate numerical value of atmospheric pressure expressed in N/m2 and as a head of water? (c) Calculate atmospheric pressure expressed as a head of mercury (the relative density of mercury is 13.6) [(c) 0.76 m] 1.3 A rectangular tank is 1.0 m long and 0.7 m wide and contains fresh water to a depth of 0.5 m. (a) What is the gauge pressure at the bottom of the tank in N/m2? (b) What is the absolute pressure at the bottom of the tank? [4905 N/m2; 105 905 N/m2] 1.4 For the tank in question 1.3, using gauge pressure, calculate (a) the mean pressure intensity on the 0.7 m wide end of the tank; (b) the mean pressure intensity on the 1.0 m long side of the tank; (c) the force on the end of the tank; and (d) the force on the side. [2453 N/m2; 2453 N/m2; 858 N; 1226 N] 1.5 A dam that retains fresh water has a vertical face. Over a one metre length of the face at the centre of the valley the water has a depth of 38 m. (a) Calculate the resultant force on this unit length of the face. (b) At what depth from the surface does the resultant force act? [7083 103 N; 25.33 m] 1.6 (a) A rectangular culvert 2.1 m wide by 1.8 m high discharges to a river channel as in Fig. 1.10. At the end of the culvert is a vertical ap gate which is hinged along its top edge, the gate having the same dimensions as the culvert. During a ood the river rises to 3.5 m above the hinge. What is the force exerted by the oodwater on the gate, and at what depth from the surface does it act? (b) A circular gate, also hinged at the top, hangs vertically at the end of a pipe discharging to the river. The gate has a radius of 0.5 m, and during a ood the hinge is 3.5 m below the water surface. What is the force exerted by the oodwater on the gate, and at what depth from the surface does it act? [(a) 163.16 103 N at 4.461 m; (b) 30.82 103 N at 4.015 m] 1.7 A gate at the end of a sewer measures 0.8 m by 1.2 m wide. It is hinged along its top edge and hangs at an angle of 30 to the vertical, this being the angle of the banks of a trapezoidal river channel. (a) Calculate the hydrostatic force on the gate and the vertical distance between the centroid of the gate, G, and the centre of pressure, P, when the river level is 0.1 m above the top of the hinge. (b) If the river level increases to 2.0 m above the hinge, what is the force and the distance GP now? (c) Has the value of GP increased or decreased, and why has it changed in this manner? [(a) 4.21 103 N, 0.090 m; (b) 22.10 103 N, 0.017 m] 1.8 A circular gate of 0.5 m radius is hinged so that it rotates about its horizontal diameter, that is it rotates about a horizontal line passing through the centroid of the gate. The gate is at the end of a pipe discharging to a river. Measured above the centroid of the gate, the head in the pipe is 6.0 m while the head in the river is 2.0 m. Assuming that the gate is initially vertical: (a) calculate the force exerted by the water in the pipe on the gate, and the distance GP between the centre of the gate, G, and the centre of pressure, P; (b) calculate the force exerted by the river water on the gate, and the distance GP; (c) by taking moments about the hinge, using the results from above, determine the net turning moment on the gate caused by the two forces acting at their respective centres of pressure on opposite sides of the gate. Explain your answer. [(a) 46 228.5 N at 0.0104 m; (b) 15 409.5 N at 0.0312 m; (c) 0 exactly, allowing for rounding errors]

34

Understanding Hydraulics

1.9 A gate which is a quarter of a circle of radius 4.0 m holds back 2.0 m of fresh water as shown in the diagram.

Fig. Q1.10

Fig. Q1.9
Calculate the magnitude and direction of the resultant hydrostatic force on a unit length of the gate. [52.05 103 N/m at 67.9 to the horizontal, acting upwards through the centre of curvature, C] 1.10 The dam in Fig. Q1.10 has a curved face, being part of a 40 m radius circle. The dam holds back water to a depth of 35 m. Calculate the magnitude and direction of the resultant hydrostatic force per metre length. [7840.6 103 N/m at 40 to the horizontal, acting downwards through the centre of curvature, C] 1.11 A 7500 tonne reinforced concrete lock structure has been constructed in a dry dock. The lock is 60 m long by 30 m wide in plan and is shaped like an open shoe box. The side walls are 8 m high. (a) Will the lock structure oat in sea water of density 1025 kg/m3, and if so, what is its draught and free-

board? (b) What additional weight will be required to sink the structure onto the sea bed if the depth of water is 5.3 m, assuming the structure is watertight? (c) If the additional weight is to be provided by a blanket of sand (density 2600 kg/m3), how thick must the layer of sand be? (1 tonne = 1000 kg). [(a) yes, 4.07 m, 3.93 m; (b) >22 352 103 N; (c) >0.5 m] 1.12 (a) Explain what is meant by a stratied uid. (b) A pressure transducer is used to measure the hydrostatic pressure on the sea bed in a tidal estuary. The water in the estuary is stratied at the point where the measurement is taken, with fresh water (1000 kg/m3) overlying saline water (1025 kg/m3). Water sampling shows that the fresh water extends from the water surface to a depth of 2.7 m. If the transducer indicates a gauge pressure of 69.73 103 N/m2, how thick is the layer of saline water? [4.3 m]

You might also like