Misplaced Pages

Settling

Article snapshot taken from Wikipedia with creative commons attribution-sharealike license. Give it a read and then ask your questions in the chat. We can research this topic together.

Settling is the process by which particulates move towards the bottom of a liquid and form a sediment . Particles that experience a force, either due to gravity or due to centrifugal motion will tend to move in a uniform manner in the direction exerted by that force. For gravity settling, this means that the particles will tend to fall to the bottom of the vessel, forming sludge or slurry at the vessel base. Settling is an important operation in many applications, such as mining , wastewater and drinking water treatment, biological science, space propellant reignition, and scooping.

#902097

84-419: For settling particles that are considered individually, i.e. dilute particle solutions, there are two main forces enacting upon any particle. The primary force is an applied force, such as gravity, and a drag force that is due to the motion of the particle through the fluid . The applied force is usually not affected by the particle's velocity, whereas the drag force is a function of the particle velocity. For

168-494: A = m d v d t = m g − 1 2 ρ v 2 A C d . {\displaystyle ma=m{\frac {\mathrm {d} v}{\mathrm {d} t}}=mg-{\frac {1}{2}}\rho v^{2}AC_{d}.} Although this is a Riccati equation that can be solved by reduction to a second-order linear differential equation, it is easier to separate variables . A more practical form of this equation can be obtained by making

252-438: A fluid ( air is the most common example). It is reached when the sum of the drag force ( F d ) and the buoyancy is equal to the downward force of gravity ( F G ) acting on the object. Since the net force on the object is zero, the object has zero acceleration . For objects falling through air at normal pressure, the buoyant force is usually dismissed and not taken into account, as its effects are negligible. As

336-476: A limit value of one, for large time t . In other words, velocity asymptotically approaches a maximum value called the terminal velocity v t : v t = 2 m g ρ A C D . {\displaystyle v_{t}={\sqrt {\frac {2mg}{\rho AC_{D}}}}.\,} For an object falling and released at relative-velocity v  = v i at time t  = 0, with v i < v t ,

420-665: A limit value of one, for large time t . Velocity asymptotically tends to the terminal velocity v t , strictly from above v t . For v i = v t , the velocity is constant: v ( t ) = v t . {\displaystyle v(t)=v_{t}.} These functions are defined by the solution of the following differential equation : g − ρ A C D 2 m v 2 = d v d t . {\displaystyle g-{\frac {\rho AC_{D}}{2m}}v^{2}={\frac {dv}{dt}}.\,} Or, more generically (where F ( v ) are

504-403: A 1920 U.S. Army Ordnance study. Competition speed skydivers fly in a head-down position and can reach speeds of 150 m/s (490 ft/s). The current record is held by Felix Baumgartner who jumped from an altitude of 38,887 m (127,582 ft) and reached 380 m/s (1,200 ft/s), though he achieved this speed at high altitude where the density of the air is much lower than at

588-526: A fluid and a solid surface. Drag forces tend to decrease fluid velocity relative to the solid object in the fluid's path. Unlike other resistive forces, drag force depends on velocity. This is because drag force is proportional to the velocity for low-speed flow and the velocity squared for high-speed flow. This distinction between low and high-speed flow is measured by the Reynolds number . Examples of drag include: Types of drag are generally divided into

672-809: A fluid at relatively slow speeds (assuming there is no turbulence). Purely laminar flow only exists up to Re = 0.1 under this definition. In this case, the force of drag is approximately proportional to velocity. The equation for viscous resistance is: F D = − b v {\displaystyle \mathbf {F} _{D}=-b\mathbf {v} \,} where: When an object falls from rest, its velocity will be v ( t ) = ( ρ − ρ 0 ) V g b ( 1 − e − b t / m ) {\displaystyle v(t)={\frac {(\rho -\rho _{0})\,V\,g}{b}}\left(1-e^{-b\,t/m}\right)} where: The velocity asymptotically approaches

756-450: A fluid increases as the cube of the velocity increases. For example, a car cruising on a highway at 50 mph (80 km/h) may require only 10 horsepower (7.5 kW) to overcome aerodynamic drag, but that same car at 100 mph (160 km/h) requires 80 hp (60 kW). With a doubling of speeds, the drag/force quadruples per the formula. Exerting 4 times the force over a fixed distance produces 4 times as much work . At twice

840-553: A fluid under its own weight can reach a terminal velocity (settling velocity) if the net force acting on the object becomes zero. When the terminal velocity is reached the weight of the object is exactly balanced by the upward buoyancy force and drag force. That is where If the falling object is spherical in shape, the expression for the three forces are given below: where Substitution of equations ( 2 – 4 ) in equation ( 1 ) and solving for terminal velocity, V t {\displaystyle V_{t}} to yield

924-693: A human body ( d {\displaystyle d} ≈0.6 m) v t {\displaystyle v_{t}} ≈70 m/s, for a small animal like a cat ( d {\displaystyle d} ≈0.2 m) v t {\displaystyle v_{t}} ≈40 m/s, for a small bird ( d {\displaystyle d} ≈0.05 m) v t {\displaystyle v_{t}} ≈20 m/s, for an insect ( d {\displaystyle d} ≈0.01 m) v t {\displaystyle v_{t}} ≈9 m/s, and so on. Terminal velocity for very small objects (pollen, etc.) at low Reynolds numbers

SECTION 10

#1732772492903

1008-401: A lower terminal velocity than one with a small projected area relative to its mass, such as a dart. In general, for the same shape and material, the terminal velocity of an object increases with size. This is because the downward force (weight) is proportional to the cube of the linear dimension, but the air resistance is approximately proportional to the cross-section area which increases only as

1092-674: A minimum at some airspeed - an aircraft flying at this speed will be at or close to its optimal efficiency. Pilots will use this speed to maximize endurance (minimum fuel consumption), or maximize gliding range in the event of an engine failure. Drag depends on the properties of the fluid and on the size, shape, and speed of the object. One way to express this is by means of the drag equation : F D = 1 2 ρ v 2 C D A {\displaystyle F_{\mathrm {D} }\,=\,{\tfrac {1}{2}}\,\rho \,v^{2}\,C_{\mathrm {D} }\,A} where The drag coefficient depends on

1176-468: A particle at rest no drag force will be exhibited, which causes the particle to accelerate due to the applied force. When the particle accelerates, the drag force acts in the direction opposite to the particle's motion, retarding further acceleration, in the absence of other forces drag directly opposes the applied force. As the particle increases in velocity eventually the drag force and the applied force will approximately equate , causing no further change in

1260-518: A settling tank with water. The oil floats to the top of the water then is collected. In drinking water and waste water treatment a flocculant or coagulant is often added prior to settling to form larger particles that settle out quickly in a settling tank or ( lamella ) clarifier , leaving the water with a lower turbidity . In winemaking , the French term for this process is débourbage . This step usually occurs in white wine production before

1344-426: A simpler form t = 1 2 α g ln ⁡ 1 + α v 1 − α v = a r t a n h ( α v ) α g , {\displaystyle t={1 \over 2\alpha g}\ln {\frac {1+\alpha v}{1-\alpha v}}={\frac {\mathrm {artanh} (\alpha v)}{\alpha g}},} with artanh

1428-468: A small sphere with radius r {\displaystyle r} = 0.5 micrometre (diameter = 1.0 μm) moving through water at a velocity v {\displaystyle v} of 10 μm/s. Using 10 Pa·s as the dynamic viscosity of water in SI units, we find a drag force of 0.09 pN. This is about the drag force that a bacterium experiences as it swims through water. The drag coefficient of

1512-623: A sphere can be determined for the general case of a laminar flow with Reynolds numbers less than 2 ⋅ 10 5 {\displaystyle 2\cdot 10^{5}} using the following formula: C D = 24 R e + 4 R e + 0.4   ;           R e < 2 ⋅ 10 5 {\displaystyle C_{D}={\frac {24}{Re}}+{\frac {4}{\sqrt {Re}}}+0.4~{\text{;}}~~~~~Re<2\cdot 10^{5}} For Reynolds numbers less than 1, Stokes' law applies and

1596-558: Is about v t = g d ρ o b j ρ . {\displaystyle v_{t}={\sqrt {gd{\frac {\rho _{obj}}{\rho }}}}.\,} For objects of water-like density (raindrops, hail, live objects—mammals, birds, insects, etc.) falling in air near Earth's surface at sea level, the terminal velocity is roughly equal to with d in metre and v t in m/s. v t = 90 d , {\displaystyle v_{t}=90{\sqrt {d}},\,} For example, for

1680-408: Is about 55 m/s (180 ft/s). This speed is the asymptotic limiting value of the speed, and the forces acting on the body balance each other more and more closely as the terminal speed is approached. In this example, a speed of 50% of terminal speed is reached after only about 3 seconds, while it takes 8 seconds to reach 90%, 15 seconds to reach 99% and so on. Higher speeds can be attained if

1764-448: Is also defined in terms of the hyperbolic tangent function: v ( t ) = v t tanh ⁡ ( t g v t + arctanh ⁡ ( v i v t ) ) . {\displaystyle v(t)=v_{t}\tanh \left(t{\frac {g}{v_{t}}}+\operatorname {arctanh} \left({\frac {v_{i}}{v_{t}}}\right)\right).\,} For v i > v t ,

SECTION 20

#1732772492903

1848-625: Is asymptotically proportional to R e − 1 {\displaystyle \mathrm {Re} ^{-1}} , which means that the drag is linearly proportional to the speed, i.e. the drag force on a small sphere moving through a viscous fluid is given by the Stokes Law : F d = 3 π μ D v {\displaystyle F_{\rm {d}}=3\pi \mu Dv} At high R e {\displaystyle \mathrm {Re} } , C D {\displaystyle C_{\rm {D}}}

1932-583: Is called Stokes' law . When the value of C d {\displaystyle C_{d}} is substituted in the equation ( 5 ), we obtain the expression for terminal speed of a spherical object moving under creeping flow conditions: V t = g d 2 18 μ ( ρ s − ρ ) , {\displaystyle V_{t}={\frac {gd^{2}}{18\mu }}\left(\rho _{s}-\rho \right),} where ρ s {\displaystyle \rho _{s}}

2016-529: Is commonly used to measure suspended solids in wastewater or stormwater runoff . The simplicity of the method makes it popular for estimating water quality . To numerically gauge the stability of suspended solids and predict agglomeration and sedimentation events, zeta potential is commonly analyzed. This parameter indicates the electrostatic repulsion between solid particles and can be used to predict whether aggregation and settling will occur over time. The water sample to be measured should be representative of

2100-399: Is determined by Stokes law. In short, terminal velocity is higher for larger creatures, and thus potentially more deadly. A creature such as a mouse falling at its terminal velocity is much more likely to survive impact with the ground than a human falling at its terminal velocity. The equation for viscous resistance or linear drag is appropriate for objects or particles moving through

2184-400: Is immediately placed in a stationary holding rack to allow quiescent settling. The rack should be located away from heating sources, including direct sunlight, which might cause currents within the cone from thermal density changes of the liquid contents. After 45 minutes of settling, the cone is partially rotated about its axis of symmetry just enough to dislodge any settled material adhering to

2268-408: Is known as bluff or blunt when the source of drag is dominated by pressure forces, and streamlined if the drag is dominated by viscous forces. For example, road vehicles are bluff bodies. For aircraft, pressure and friction drag are included in the definition of parasitic drag . Parasite drag is often expressed in terms of a hypothetical. This is the area of a flat plate perpendicular to the flow. It

2352-406: Is made up of multiple components including viscous pressure drag ( form drag ), and drag due to surface roughness ( skin friction drag ). Additionally, the presence of multiple bodies in relative proximity may incur so called interference drag , which is sometimes described as a component of parasitic drag. In aviation, induced drag tends to be greater at lower speeds because a high angle of attack

2436-428: Is more or less constant, but drag will vary as the square of the speed varies. The graph to the right shows how C D {\displaystyle C_{\rm {D}}} varies with R e {\displaystyle \mathrm {Re} } for the case of a sphere. Since the power needed to overcome the drag force is the product of the force times speed, the power needed to overcome drag will vary as

2520-472: Is presented at Drag equation § Derivation . The reference area A is often the orthographic projection of the object, or the frontal area, on a plane perpendicular to the direction of motion. For objects with a simple shape, such as a sphere, this is the cross sectional area. Sometimes a body is a composite of different parts, each with a different reference area (drag coefficient corresponding to each of those different areas must be determined). In

2604-473: Is required to maintain lift, creating more drag. However, as speed increases the angle of attack can be reduced and the induced drag decreases. Parasitic drag, however, increases because the fluid is flowing more quickly around protruding objects increasing friction or drag. At even higher speeds ( transonic ), wave drag enters the picture. Each of these forms of drag changes in proportion to the others based on speed. The combined overall drag curve therefore shows

Settling - Misplaced Pages Continue

2688-405: Is that of Schiller and Naumann, and may be valid for 0.2 ≤ R e ≤ 1000 {\displaystyle 0.2\leq Re\leq 1000} : Stokes, transitional and Newtonian settling describe the behaviour of a single spherical particle in an infinite fluid, known as free settling. However this model has limitations in practical application. Alternate considerations, such as

2772-583: Is the Reynolds number related to fluid path length L. As mentioned, the drag equation with a constant drag coefficient gives the force moving through fluid a relatively large velocity, i.e. high Reynolds number , Re > ~1000. This is also called quadratic drag . F D = 1 2 ρ v 2 C D A , {\displaystyle F_{D}\,=\,{\tfrac {1}{2}}\,\rho \,v^{2}\,C_{D}\,A,} The derivation of this equation

2856-447: Is the density of the object. The creeping flow results can be applied in order to study the settling of sediments near the ocean bottom and the fall of moisture drops in the atmosphere. The principle is also applied in the falling sphere viscometer , an experimental device used to measure the viscosity of highly viscous fluids, for example oil, paraffin, tar etc. When the buoyancy effects are taken into account, an object falling through

2940-412: Is the induced drag. Another drag component, namely wave drag , D w {\displaystyle D_{w}} , results from shock waves in transonic and supersonic flight speeds. The shock waves induce changes in the boundary layer and pressure distribution over the body surface. Terminal velocity Terminal velocity is the maximum speed attainable by an object as it falls through

3024-755: Is the wind speed and v o {\displaystyle v_{o}} is the object speed (both relative to ground). Velocity as a function of time for an object falling through a non-dense medium, and released at zero relative-velocity v  = 0 at time t  = 0, is roughly given by a function involving a hyperbolic tangent (tanh): v ( t ) = 2 m g ρ A C D tanh ⁡ ( t g ρ C D A 2 m ) . {\displaystyle v(t)={\sqrt {\frac {2mg}{\rho AC_{D}}}}\tanh \left(t{\sqrt {\frac {g\rho C_{D}A}{2m}}}\right).\,} The hyperbolic tangent has

3108-547: Is used when comparing the drag of different aircraft For example, the Douglas DC-3 has an equivalent parasite area of 2.20 m (23.7 sq ft) and the McDonnell Douglas DC-9 , with 30 years of advancement in aircraft design, an area of 1.91 m (20.6 sq ft) although it carried five times as many passengers. Lift-induced drag (also called induced drag ) is drag which occurs as

3192-728: The inverse hyperbolic tangent function. Alternatively, 1 α tanh ⁡ ( α g t ) = v , {\displaystyle {\frac {1}{\alpha }}\tanh(\alpha gt)=v,} with tanh the hyperbolic tangent function. Assuming that g is positive (which it was defined to be), and substituting α back in, the speed v becomes v = 2 m g ρ A C d tanh ⁡ ( t g ρ A C d 2 m ) . {\displaystyle v={\sqrt {\frac {2mg}{\rho AC_{d}}}}\tanh \left(t{\sqrt {\frac {g\rho AC_{d}}{2m}}}\right).} Using

3276-410: The lift coefficient also increases, and so too does the lift-induced drag. At the onset of stall , lift is abruptly decreased, as is lift-induced drag, but viscous pressure drag, a component of parasite drag, increases due to the formation of turbulent unattached flow in the wake behind the body. Parasitic drag , or profile drag, is drag caused by moving a solid object through a fluid. Parasitic drag

3360-413: The order 10 ). For an object with well-defined fixed separation points, like a circular disk with its plane normal to the flow direction, the drag coefficient is constant for Re  > 3,500. The further the drag coefficient C d is, in general, a function of the orientation of the flow with respect to the object (apart from symmetrical objects like a sphere). Under the assumption that

3444-467: The terminal velocity of a particle in a Newtonian regime can again be obtained by equating the drag force to the applied force, resulting in the following expression In the intermediate region between Stokes drag and Newtonian drag, there exists a transitional regime, where the analytical solution to the problem of a falling sphere becomes problematic. To solve this, empirical expressions are used to calculate drag in this region. One such empirical equation

Settling - Misplaced Pages Continue

3528-465: The Earth's surface, producing a correspondingly lower drag force. The biologist J. B. S. Haldane wrote, To the mouse and any smaller animal [gravity] presents practically no dangers. You can drop a mouse down a thousand-yard mine shaft; and, on arriving at the bottom, it gets a slight shock and walks away. A rat is killed, a man is broken, a horse splashes. For the resistance presented to movement by

3612-405: The air is proportional to the surface of the moving object. Divide an animal's length, breadth, and height each by ten; its weight is reduced to a thousandth, but its surface only to a hundredth. So the resistance to falling in the case of the small animal is relatively ten times greater than the driving force. For terminal velocity in falling through air, where viscosity is negligible compared to

3696-413: The airflow and forces the flow to move downward. This results in an equal and opposite force acting upward on the wing which is the lift force. The change of momentum of the airflow downward results in a reduction of the rearward momentum of the flow which is the result of a force acting forward on the airflow and applied by the wing to the air flow; an equal but opposite force acts on the wing rearward which

3780-407: The airplane produces lift, another drag component results. Induced drag , symbolized D i {\displaystyle D_{i}} , is due to a modification of the pressure distribution due to the trailing vortex system that accompanies the lift production. An alternative perspective on lift and drag is gained from considering the change of momentum of the airflow. The wing intercepts

3864-453: The case of a wing , the reference areas are the same, and the drag force is in the same ratio as the lift force . Therefore, the reference for a wing is often the lifting area, sometimes referred to as "wing area" rather than the frontal area. For an object with a smooth surface, and non-fixed separation points (like a sphere or circular cylinder), the drag coefficient may vary with Reynolds number Re , up to extremely high values ( Re of

3948-426: The creeping flow around a sphere was first given by Stokes in 1851. From Stokes' solution, the drag force acting on the sphere of diameter d {\displaystyle d} can be obtained as where the Reynolds number, R e = ρ d μ V {\displaystyle Re={\frac {\rho d}{\mu }}V} . The expression for the drag force given by equation ( 6 )

4032-678: The drag coefficient C D {\displaystyle C_{\rm {D}}} as a function of Bejan number and the ratio between wet area A w {\displaystyle A_{\rm {w}}} and front area A f {\displaystyle A_{\rm {f}}} : C D = 2 A w A f B e R e L 2 {\displaystyle C_{\rm {D}}=2{\frac {A_{\rm {w}}}{A_{\rm {f}}}}{\frac {\mathrm {Be} }{\mathrm {Re} _{L}^{2}}}} where R e L {\displaystyle \mathrm {Re} _{L}}

4116-487: The drag coefficient approaches 24 R e {\displaystyle {\frac {24}{Re}}} ! In aerodynamics , aerodynamic drag , also known as air resistance , is the fluid drag force that acts on any moving solid body in the direction of the air's freestream flow. Alternatively, calculated from the flow field perspective (far-field approach), the drag force results from three natural phenomena: shock waves , vortex sheet, and viscosity . When

4200-620: The drag constant: b = 6 π η r {\displaystyle b=6\pi \eta r\,} where r {\displaystyle r} is the Stokes radius of the particle, and η {\displaystyle \eta } is the fluid viscosity. The resulting expression for the drag is known as Stokes' drag : F D = − 6 π η r v . {\displaystyle \mathbf {F} _{D}=-6\pi \eta r\,\mathbf {v} .} For example, consider

4284-401: The drag force, and without considering buoyancy effects, terminal velocity is given by V t = 2 m g ρ A C d {\displaystyle V_{t}={\sqrt {\frac {2mg}{\rho AC_{d}}}}} where In reality, an object approaches its terminal speed asymptotically . Buoyancy effects, due to the upward force on the object by

SECTION 50

#1732772492903

4368-412: The drag rule is not simple as a single sphere settling in a stationary fluid. However, this knowledge indicates how drag behaves in more complex systems, which are designed and studied by engineers applying empirical and more sophisticated tools. For example, 'settling tanks ' are used for separating solids and/or oil from another liquid. In food processing , the vegetable is crushed and placed inside of

4452-483: The fall, speed decreases to change with the local terminal speed. Using mathematical terms, defining down to be positive, the net force acting on an object falling near the surface of Earth is (according to the drag equation ): F net = m a = m g − 1 2 ρ v 2 A C d , {\displaystyle F_{\text{net}}=ma=mg-{\frac {1}{2}}\rho v^{2}AC_{d},} with v ( t )

4536-613: The fluid are negligible (assumption of massless fluid) in comparison to other forces. Such flows are called creeping or Stokes flows and the condition to be satisfied for the flows to be creeping flows is the Reynolds number , R e ≪ 1 {\displaystyle Re\ll 1} . The equation of motion for creeping flow (simplified Navier–Stokes equation ) is given by: ∇ p = μ ∇ 2 v {\displaystyle {\mathbf {\nabla } }p=\mu \nabla ^{2}{\mathbf {v} }} where: The analytical solution for

4620-441: The fluid is not moving relative to the currently used reference system, the power required to overcome the aerodynamic drag is given by: P D = F D ⋅ v = 1 2 ρ v 3 A C D {\displaystyle P_{D}=\mathbf {F} _{D}\cdot \mathbf {v} ={\tfrac {1}{2}}\rho v^{3}AC_{D}} The power needed to push an object through

4704-411: The following categories: The effect of streamlining on the relative proportions of skin friction and form drag is shown for two different body sections: An airfoil, which is a streamlined body, and a cylinder, which is a bluff body. Also shown is a flat plate illustrating the effect that orientation has on the relative proportions of skin friction, and pressure difference between front and back. A body

4788-448: The forces acting on the object beyond drag): 1 m ∑ F ( v ) − ρ A C D 2 m v 2 = d v d t . {\displaystyle {\frac {1}{m}}\sum F(v)-{\frac {\rho AC_{D}}{2m}}v^{2}={\frac {dv}{dt}}.\,} For a potato-shaped object of average diameter d and of density ρ obj , terminal velocity

4872-472: The formula for terminal velocity V t = 2 m g ρ A C d {\displaystyle V_{t}={\sqrt {\frac {2mg}{\rho AC_{d}}}}} the equation can be rewritten as v = V t tanh ⁡ ( t g V t ) . {\displaystyle v=V_{t}\tanh \left(t{\frac {g}{V_{t}}}\right).} As time tends to infinity ( t → ∞),

4956-400: The hyperbolic tangent tends to 1, resulting in the terminal speed V t = lim t → ∞ v ( t ) = 2 m g ρ A C d . {\displaystyle V_{t}=\lim _{t\to \infty }v(t)={\sqrt {\frac {2mg}{\rho AC_{d}}}}.} For very slow motion of the fluid, the inertia forces of

5040-427: The increasing importance of fluid inertia, requiring the use of empirical solutions to calculate drag forces. Defining a drag coefficient , C d {\displaystyle C_{d}} , as the ratio of the force experienced by the particle divided by the impact pressure of the fluid, a coefficient that can be considered as the transfer of available fluid force into drag is established. In this region

5124-534: The inertia of the impacting fluid is responsible for the majority of force transfer to the particle. For a spherical particle in the Stokes regime this value is not constant, however in the Newtonian drag regime the drag on a sphere can be approximated by a constant, 0.44. This constant value implies that the efficiency of transfer of energy from the fluid to the particle is not a function of fluid velocity. As such

SECTION 60

#1732772492903

5208-1395: The integral of both sides yields ∫ 0 t d t ′ = 1 g ∫ 0 v d v ′ 1 − α 2 v ′ 2 . {\displaystyle \int _{0}^{t}{\mathrm {d} t'}={1 \over g}\int _{0}^{v}{\frac {\mathrm {d} v'}{1-\alpha ^{2}v^{\prime 2}}}.} After integration, this becomes t − 0 = 1 g [ ln ⁡ ( 1 + α v ′ ) 2 α − ln ⁡ ( 1 − α v ′ ) 2 α + C ] v ′ = 0 v ′ = v = 1 g [ ln ⁡ 1 + α v ′ 1 − α v ′ 2 α + C ] v ′ = 0 v ′ = v {\displaystyle t-0={1 \over g}\left[{\ln(1+\alpha v') \over 2\alpha }-{\frac {\ln(1-\alpha v')}{2\alpha }}+C\right]_{v'=0}^{v'=v}={1 \over g}\left[{\ln {\frac {1+\alpha v'}{1-\alpha v'}} \over 2\alpha }+C\right]_{v'=0}^{v'=v}} or in

5292-541: The interaction of particles in the fluid, or the interaction of the particles with the container walls can modify the settling behaviour. Settling that has these forces in appreciable magnitude is known as hindered settling. Subsequently, semi-analytic or empirical solutions may be used to perform meaningful hindered settling calculations. The solid-gas flow systems are present in many industrial applications, as dry, catalytic reactors, settling tanks, pneumatic conveying of solids, among others. Obviously, in industrial operations

5376-773: The natural sciences, and is given by: where w is the settling velocity, ρ is density (the subscripts p and f indicate particle and fluid respectively), g is the acceleration due to gravity, r is the radius of the particle and μ is the dynamic viscosity of the fluid. Stokes' law applies when the Reynolds number , Re, of the particle is less than 0.1. Experimentally Stokes' law is found to hold within 1% for R e ≤ 0.1 {\displaystyle Re\leq 0.1} , within 3% for R e ≤ 0.5 {\displaystyle Re\leq 0.5} and within 9% R e ≤ 1.0 {\displaystyle Re\leq 1.0} . With increasing Reynolds numbers, Stokes law begins to break down due to

5460-409: The particle's velocity. This velocity is known as the terminal velocity , settling velocity or fall velocity of the particle. This is readily measurable by examining the rate of fall of individual particles. The terminal velocity of the particle is affected by many parameters, i.e. anything that will alter the particle's drag. Hence the terminal velocity is most notably dependent upon grain size ,

5544-401: The properties of the fluid, the mass of the object and its projected cross-sectional surface area . Air density increases with decreasing altitude, at about 1% per 80 metres (260 ft) (see barometric formula ). For objects falling through the atmosphere, for every 160 metres (520 ft) of fall, the terminal speed decreases 1%. After reaching the local terminal velocity, while continuing

5628-404: The result of the creation of lift on a three-dimensional lifting body , such as the wing or propeller of an airplane. Induced drag consists primarily of two components: drag due to the creation of trailing vortices ( vortex drag ); and the presence of additional viscous drag ( lift-induced viscous drag ) that is not present when lift is zero. The trailing vortices in the flow-field, present in

5712-414: The shape (roundness and sphericity) and density of the grains, as well as to the viscosity and density of the fluid. For dilute suspensions, Stokes' law predicts the settling velocity of small spheres in fluid , either air or water. This originates due to the strength of viscous forces at the surface of the particle providing the majority of the retarding force. Stokes' law finds many applications in

5796-408: The shape of the object and on the Reynolds number R e = v D ν = ρ v D μ , {\displaystyle \mathrm {Re} ={\frac {vD}{\nu }}={\frac {\rho vD}{\mu }},} where At low R e {\displaystyle \mathrm {Re} } , C D {\displaystyle C_{\rm {D}}}

5880-420: The side of the cone. Accumulated sediment is observed and measured fifteen minutes later, after one hour of total settling time. Drag (physics) In fluid dynamics , drag , sometimes referred to as fluid resistance , is a force acting opposite to the relative motion of any object moving with respect to a surrounding fluid . This can exist between two fluid layers, two solid surfaces, or between

5964-411: The skydiver pulls in their limbs (see also freeflying ). In this case, the terminal speed increases to about 90 m/s (300 ft/s), which is almost the terminal speed of the peregrine falcon diving down on its prey. The same terminal speed is reached for a typical .30-06 bullet dropping downwards—when it is returning to the ground having been fired upwards, or dropped from a tower—according to

6048-443: The speed of an object increases, so does the drag force acting on it, which also depends on the substance it is passing through (for example air or water). At some speed, the drag or force of resistance will equal the gravitational pull on the object. At this point the object stops accelerating and continues falling at a constant speed called the terminal velocity (also called settling velocity ). An object moving downward faster than

6132-808: The speed, the work (resulting in displacement over a fixed distance) is done twice as fast. Since power is the rate of doing work, 4 times the work done in half the time requires 8 times the power. When the fluid is moving relative to the reference system, for example, a car driving into headwind, the power required to overcome the aerodynamic drag is given by the following formula: P D = F D ⋅ v o = 1 2 C D A ρ ( v w + v o ) 2 v o {\displaystyle P_{D}=\mathbf {F} _{D}\cdot \mathbf {v_{o}} ={\tfrac {1}{2}}C_{D}A\rho (v_{w}+v_{o})^{2}v_{o}} Where v w {\displaystyle v_{w}}

6216-443: The square of the linear dimension. For very small objects such as dust and mist, the terminal velocity is easily overcome by convection currents which can prevent them from reaching the ground at all, and hence they can stay suspended in the air for indefinite periods. Air pollution and fog are examples. Based on air resistance, for example, the terminal speed of a skydiver in a belly-to-earth (i.e., face down) free fall position

6300-884: The square of the speed at low Reynolds numbers, and as the cube of the speed at high numbers. It can be demonstrated that drag force can be expressed as a function of a dimensionless number, which is dimensionally identical to the Bejan number . Consequently, drag force and drag coefficient can be a function of Bejan number. In fact, from the expression of drag force it has been obtained: F d = Δ p A w = 1 2 C D A f ν μ l 2 R e L 2 {\displaystyle F_{\rm {d}}=\Delta _{\rm {p}}A_{\rm {w}}={\frac {1}{2}}C_{\rm {D}}A_{\rm {f}}{\frac {\nu \mu }{l^{2}}}\mathrm {Re} _{L}^{2}} and consequently allows expressing

6384-442: The start of fermentation . Settleable solids are the particulates that settle out of a still fluid. Settleable solids can be quantified for a suspension using an Imhoff cone. The standard Imhoff cone of transparent glass or plastic holds one liter of liquid and has calibrated markings to measure the volume of solids accumulated in the bottom of the conical container after settling for one hour. A standardized Imhoff cone procedure

6468-633: The substitution α = ⁠ ρAC d / 2 mg ⁠ . Dividing both sides by m gives d v d t = g ( 1 − α 2 v 2 ) . {\displaystyle {\frac {\mathrm {d} v}{\mathrm {d} t}}=g\left(1-\alpha ^{2}v^{2}\right).} The equation can be re-arranged into d t = d v g ( 1 − α 2 v 2 ) . {\displaystyle \mathrm {d} t={\frac {\mathrm {d} v}{g(1-\alpha ^{2}v^{2})}}.} Taking

6552-600: The surrounding fluid, can be taken into account using Archimedes' principle : the mass m {\displaystyle m} has to be reduced by the displaced fluid mass ρ V {\displaystyle \rho V} , with V {\displaystyle V} the volume of the object. So instead of m {\displaystyle m} use the reduced mass m r = m − ρ V {\displaystyle m_{r}=m-\rho V} in this and subsequent formulas. The terminal speed of an object changes due to

6636-482: The terminal velocity v t = ( ρ − ρ 0 ) V g b {\displaystyle v_{t}={\frac {(\rho -\rho _{0})Vg}{b}}} . For a given b {\displaystyle b} , denser objects fall more quickly. For the special case of small spherical objects moving slowly through a viscous fluid (and thus at small Reynolds number), George Gabriel Stokes derived an expression for

6720-399: The terminal velocity (for example because it was thrown downwards, it fell from a thinner part of the atmosphere, or it changed shape) will slow down until it reaches the terminal velocity. Drag depends on the projected area , here represented by the object's cross-section or silhouette in a horizontal plane. An object with a large projected area relative to its mass, such as a parachute, has

6804-413: The total stream. Samples are best collected from the discharge falling from a pipe or over a weir, because samples skimmed from the top of a flowing channel may fail to capture larger, high-density solids moving along the bottom of the channel. The sampling bucket is vigorously stirred to uniformly re-suspend all collected solids immediately before pouring the volume required to fill the cone. The filled cone

6888-489: The velocity function is defined in terms of the hyperbolic cotangent function: v ( t ) = v t coth ⁡ ( t g v t + coth − 1 ⁡ ( v i v t ) ) . {\displaystyle v(t)=v_{t}\coth \left(t{\frac {g}{v_{t}}}+\coth ^{-1}\left({\frac {v_{i}}{v_{t}}}\right)\right).\,} The hyperbolic cotangent also has

6972-512: The velocity of the object as a function of time t . At equilibrium , the net force is zero ( F net = 0) and the velocity becomes the terminal velocity lim t →∞ v ( t ) = V t : m g − 1 2 ρ V t 2 A C d = 0. {\displaystyle mg-{1 \over 2}\rho V_{t}^{2}AC_{d}=0.} Solving for V t yields: The drag equation is—assuming ρ , g and C d to be constants: m

7056-405: The wake of a lifting body, derive from the turbulent mixing of air from above and below the body which flows in slightly different directions as a consequence of creation of lift . With other parameters remaining the same, as the lift generated by a body increases, so does the lift-induced drag. This means that as the wing's angle of attack increases (up to a maximum called the stalling angle),

#902097