NTU method

Last updated

The number of transfer units (NTU) method is used to calculate the rate of heat transfer in heat exchangers (especially parallel flow, counter current, and cross-flow exchangers) when there is insufficient information to calculate the log mean temperature difference (LMTD). Alternatively, this method is useful for determining the expected heat exchanger effectiveness from the known geometry. In heat exchanger analysis, if the fluid inlet and outlet temperatures are specified or can be determined by simple energy balance, the LMTD method can be used; but when these temperatures are not available either the NTU or the effectiveness NTU method is used.

Contents

The effectiveness-NTU method is very useful for all the flow arrangements (besides parallel flow, cross flow, and counterflow ones) but the effectiveness of all other types must be obtained by a numerical solution of the partial differential equations and there is no analytical equation for LMTD or effectiveness.

Defining and using heat exchanger effectiveness

To define the effectiveness of a heat exchanger we need to find the maximum possible heat transfer that can be hypothetically achieved in a counter-flow heat exchanger of infinite length. Therefore one fluid will experience the maximum possible temperature difference, which is the difference of (the temperature difference between the inlet temperature of the hot stream and the inlet temperature of the cold stream). First, you must know the specific heat capacity of your two air streams, denoted as . By definition is the derivative of enthalpy with respect to temperature:

This information can usually be found in a thermodynamics textbook, [1] or by using various software packages. Additionally, the mass flowrates () of the two streams exchanging heat must be known (here, the cold stream is denoted with subscripts 'c' and the hot stream is denoted with subscripts 'h'). The method proceeds by calculating the heat capacity rates (i.e. mass flow rate multiplied by specific heat capacity) and for the hot and cold fluids respectively. To determine the maximum possible heat transfer rate in the heat exchanger, the minimum heat capacity rate must be used, denoted as :

Where is the mass flow rate and is the fluid's specific heat capacity at constant pressure. The maximum possible heat transfer rate is then determined by the following expression:

Here, is the maximum rate of heat that could be transferred between the fluids per unit time. must be used as it is the fluid with the lowest heat capacity rate that would, in this hypothetical infinite length exchanger, actually undergo the maximum possible temperature change. The other fluid would change temperature more slowly along the heat exchanger length. The method, at this point, is concerned only with the fluid undergoing the maximum temperature change.

The effectiveness of the heat exchanger (), is the ratio between the actual heat transfer rate and the maximum possible heat transfer rate:

where the real heat transfer rate can be determined either from the cold fluid or the hot fluid (they must provide equivalent results):

Effectiveness is a dimensionless quantity between 0 and 1. If we know for a particular heat exchanger, and we know the inlet conditions of the two flow streams we can calculate the amount of heat being transferred between the fluids by:

Then, having determined the actual heat transfer from the effectiveness and inlet temperatures, the outlet temperatures can be determined from the equation above.

Relating Effectiveness to the Number of Transfer Units (NTU)

For any heat exchanger it can be shown that the effectiveness of the heat exchanger is related to a non-dimensional term called the "number of transfer units" or NTU:

For a given geometry, can be calculated using correlations in terms of the "heat capacity ratio," or and NTU:

describes heat transfer across a surface

Here, is the overall heat transfer coefficient, is the total heat transfer area, and is the minimum heat capacity rate. To better understand where this definition of NTU comes from, consider the follow heat transfer energy balance, which is an extension of the energy balance above:

From this energy balance, it is clear that NTU relates the temperature change of the flow with the minimum heat capacitance rate to the log mean temperature difference (). Starting from the differential equations that describe heat transfer, several "simple" correlations between effectiveness and NTU can be made. [2] For brevity, below summarizes the Effectiveness-NTU correlations for some of the most common flow configurations:

For example, the effectiveness of a parallel flow heat exchanger is calculated with:

Or the effectiveness of a counter-current flow heat exchanger is calculated with:

For a balanced counter-current flow heat exchanger (balanced meaning , which is a scenario desirable to reduce entropy):

A single-stream heat exchanger is a special case in which . This occurs when or and may represent a situation in which a phase change (condensation or evaporation) is occurring in one of the heat exchanger fluids or when one of the heat exchanger fluids is being held at a fixed temperature. In this special case the heat exchanger behavior is independent of the flow arrangement and the effectiveness is given by: [3]

For a crossflow heat exchanger with both fluid unmixed, the effectiveness is:

where is the polynomial function

If both fluids are mixed in the crossflow heat exchanger, then

If one of the fluids in the crossflow heat exchanger is mixed and the other is unmixed, the result depends on which one has the minimum heat capacity rate. If corresponds to the mixed fluid, the result is

whereas if corresponds to the unmixed fluid, the solution is

All these formulas for crossflow heat exchangers are also valid for .

Additional effectiveness-NTU analytical relationships have been derived for other flow arrangements, including shell-and-tube heat exchangers with multiple passes and different shell types, and plate heat exchangers. [4]


Effectiveness-NTU method for gaseous mass transfer

It is common in the field of mass transfer system design and modeling to draw analogies between heat transfer and mass transfer. [2] However, a mass transfer-analogous definition of the effectiveness-NTU method requires some additional terms. One common misconception is that gaseous mass transfer is driven by concentration gradients, however, in reality it is the partial pressure of the given gas that drive mass transfer. In the same way that the heat transfer definition includes the specific heat capacity of the fluid, which describes the change in enthalpy of the fluid with respect to change in temperature and is defined as:

then a mass transfer-analogous specific mass capacity is required. This specific mass capacity should describe the change in concentration of the transferring gas relative to the partial pressure difference driving the mass transfer. This results in a definition for specific mass capacity as follows:

Here, represents the mass ratio of gas 'x' (meaning mass of gas 'x' relative to the mass of all other non-'x' gas mass) and is the partial pressure of gas 'x'. Using the ideal gas formulation for the mass ratio gives the following definition for the specific mass capacity:

Here, is the molecular weight of gas 'x' and is the average molecular weight of all other gas constituents. With this information, the NTU for gaseous mass transfer of gas 'x' can be defined as follows:

Here, is the overall mass transfer coefficient, which could be determined by empirical correlations, is the surface area for mass transfer (particularly relevant in membrane-based separations), and is the mass flowrate of bulk fluid (e.g., mass flowrate of air in an application where water vapor is being separated from the air mixture). At this point, all of the same heat transfer effectiveness-NTU correlations will accurately predict the mass transfer performance, as long as the heat transfer terms in the definition of NTU have been replaced by the mass transfer terms, as shown above. Similarly, it follows that the definition of becomes:

Effectiveness-NTU method for dehumidification applications

One particularly useful application for the above described effectiveness-NTU framework is membrane-based air dehumidification. [5] In this case, the definition of specific mass capacity can be defined for humid air and is termed "specific humidity capacity." [2]

Here, is the molecular weight of water (vapor), is the average molecular weight of air, is the partial pressure of air (not including the partial pressure of water vapor in an air mixture) and can be approximated by knowing the partial pressure of water vapor at the inlet, before dehumidification occurs, . From here, all of the previously described equations can be used to determine the effectiveness of the mass exchanger.

Importance of defining the specific mass capacity

It is very common, especially in dehumidification applications, to define the mass transfer driving force as the concentration difference. When deriving effectiveness-NTU correlations for membrane-based gas separations, this is valid only if the total pressures are approximately equal on both sides of the membrane (e.g., an energy recovery ventilator for a building). This is sufficient since the partial pressure and concentration are proportional. However, if the total pressures are not approximately equal on both sides of the membrane, the low pressure side could have a higher "concentration" but a lower partial pressure of the given gas (e.g., water vapor in a dehumidification application) than the high pressure side, thus using the concentration as the driving is not physically accurate.

Related Research Articles

<span class="mw-page-title-main">Enthalpy</span> Measure of energy in a thermodynamic system

In thermodynamics, enthalpy, is the sum of a thermodynamic system's internal energy and the product of its pressure and volume. It is a state function used in many measurements in chemical, biological, and physical systems at a constant external pressure, which is conveniently provided by the large ambient atmosphere. The pressure–volume term expresses the work that was done against constant external pressure to establish the system's physical dimensions from to some final volume , i.e. to make room for it by displacing its surroundings. The pressure-volume term is very small for solids and liquids at common conditions, and fairly small for gases. Therefore, enthalpy is a stand-in for energy in chemical systems; bond, lattice, solvation, and other chemical "energies" are actually enthalpy differences. As a state function, enthalpy depends only on the final configuration of internal energy, pressure, and volume, not on the path taken to achieve it.

In thermal fluid dynamics, the Nusselt number is the ratio of convective to conductive heat transfer at a boundary in a fluid. Convection includes both advection and diffusion (conduction). The conductive component is measured under the same conditions as the convective but for a hypothetically motionless fluid. It is a dimensionless number, closely related to the fluid's Rayleigh number.

In fluid mechanics, the Grashof number is a dimensionless number which approximates the ratio of the buoyancy to viscous forces acting on a fluid. It frequently arises in the study of situations involving natural convection and is analogous to the Reynolds number.

<span class="mw-page-title-main">Helmholtz free energy</span> Thermodynamic potential

In thermodynamics, the Helmholtz free energy is a thermodynamic potential that measures the useful work obtainable from a closed thermodynamic system at a constant temperature (isothermal). The change in the Helmholtz energy during a process is equal to the maximum amount of work that the system can perform in a thermodynamic process in which temperature is held constant. At constant temperature, the Helmholtz free energy is minimized at equilibrium.

<span class="mw-page-title-main">Internal energy</span> Energy contained within a system

The internal energy of a thermodynamic system is the energy contained within it, measured as the quantity of energy necessary to bring the system from its standard internal state to its present internal state of interest, accounting for the gains and losses of energy due to changes in its internal state, including such quantities as magnetization. It excludes the kinetic energy of motion of the system as a whole and the potential energy of position of the system as a whole, with respect to its surroundings and external force fields. It includes the thermal energy, i.e., the constituent particles' kinetic energies of motion relative to the motion of the system as a whole. The internal energy of an isolated system cannot change, as expressed in the law of conservation of energy, a foundation of the first law of thermodynamics.

<span class="mw-page-title-main">Rankine–Hugoniot conditions</span> Concept in physics

The Rankine–Hugoniot conditions, also referred to as Rankine–Hugoniot jump conditions or Rankine–Hugoniot relations, describe the relationship between the states on both sides of a shock wave or a combustion wave in a one-dimensional flow in fluids or a one-dimensional deformation in solids. They are named in recognition of the work carried out by Scottish engineer and physicist William John Macquorn Rankine and French engineer Pierre Henri Hugoniot.

In quantum mechanics, the canonical commutation relation is the fundamental relation between canonical conjugate quantities. For example,

In thermodynamics, the heat transfer coefficient or film coefficient, or film effectiveness, is the proportionality constant between the heat flux and the thermodynamic driving force for the flow of heat. It is used in calculating the heat transfer, typically by convection or phase transition between a fluid and a solid. The heat transfer coefficient has SI units in watts per square meter per kelvin (W/m²K).

There are two different Bejan numbers (Be) used in the scientific domains of thermodynamics and fluid mechanics. Bejan numbers are named after Adrian Bejan.

The Stanton number, St, is a dimensionless number that measures the ratio of heat transferred into a fluid to the thermal capacity of fluid. The Stanton number is named after Thomas Stanton (engineer) (1865–1931). It is used to characterize heat transfer in forced convection flows.

The Franz–Keldysh effect is a change in optical absorption by a semiconductor when an electric field is applied. The effect is named after the German physicist Walter Franz and Russian physicist Leonid Keldysh.

<span class="mw-page-title-main">Genus of a multiplicative sequence</span> A ring homomorphism from the cobordism ring of manifolds to another ring

In mathematics, a genus of a multiplicative sequence is a ring homomorphism from the ring of smooth compact manifolds up to the equivalence of bounding a smooth manifold with boundary to another ring, usually the rational numbers, having the property that they are constructed from a sequence of polynomials in characteristic classes that arise as coefficients in formal power series with good multiplicative properties.

The turbulent Prandtl number (Prt) is a non-dimensional term defined as the ratio between the momentum eddy diffusivity and the heat transfer eddy diffusivity. It is useful for solving the heat transfer problem of turbulent boundary layer flows. The simplest model for Prt is the Reynolds analogy, which yields a turbulent Prandtl number of 1. From experimental data, Prt has an average value of 0.85, but ranges from 0.7 to 0.9 depending on the Prandtl number of the fluid in question.

Concentric Tube Heat Exchangers are used in a variety of industries for purposes such as material processing, food preparation, and air-conditioning. They create a temperature driving force by passing fluid streams of different temperatures parallel to each other, separated by a physical boundary in the form of a pipe. This induces forced convection, transferring heat to/from the product.

Slip ratio in gas–liquid (two-phase) flow, is defined as the ratio of the velocity of the gas phase to the velocity of the liquid phase.

Heat transfer physics describes the kinetics of energy storage, transport, and energy transformation by principal energy carriers: phonons, electrons, fluid particles, and photons. Heat is energy stored in temperature-dependent motion of particles including electrons, atomic nuclei, individual atoms, and molecules. Heat is transferred to and from matter by the principal energy carriers. The state of energy stored within matter, or transported by the carriers, is described by a combination of classical and quantum statistical mechanics. The energy is different made (converted) among various carriers. The heat transfer processes are governed by the rates at which various related physical phenomena occur, such as the rate of particle collisions in classical mechanics. These various states and kinetics determine the heat transfer, i.e., the net rate of energy storage or transport. Governing these process from the atomic level to macroscale are the laws of thermodynamics, including conservation of energy.

In optics, the Ewald–Oseen extinction theorem, sometimes referred to as just the extinction theorem, is a theorem that underlies the common understanding of scattering. It is named after Paul Peter Ewald and Carl Wilhelm Oseen, who proved the theorem in crystalline and isotropic media, respectively, in 1916 and 1915. Originally, the theorem applied to scattering by an isotropic dielectric objects in free space. The scope of the theorem was greatly extended to encompass a wide variety of bianisotropic media.

The Frenkel–Kontorova model, also known as the FK model, is a fundamental model of low-dimensional nonlinear physics.

Adding controlled noise from predetermined distributions is a way of designing differentially private mechanisms. This technique is useful for designing private mechanisms for real-valued functions on sensitive data. Some commonly used distributions for adding noise include Laplace and Gaussian distributions.

Specific humidity capacity relates changes in humidity content in air (ω) to changes in vapor partial pressure and is defined as . It is analogous to specific heat capacity used in heat transfer that relates changes in enthalpy to changes in temperature (dh/dT). This thermodynamic property of air is useful when using correlations for membrane based dehumidification mass transfer models. The introduction of the specific humidity capacity allows the definition of NTU for dehumidification to be entirely analogous to the definition of NTU for heat transfer since it incorporated the vapor partial pressure as the driving force for mass transfer. Using the specific humidity capacity is required when dealing with any form of vacuum membrane dehumidification, where humidity concentration difference across the membrane is not a suitable representation of the mass transfer driving force.

References

  1. Çengel, Yunus A.; Boles, Michael A. (2015). Thermodynamics: an engineering approach (8. edition in SI units ed.). New York, NY: McGraw-Hill Education. ISBN   978-0-07-339817-4.
  2. 1 2 3 Fix, Andrew J.; Braun, James E.; Warsinger, David M. (2024-01-05). "A general effectiveness-NTU modeling framework for membrane dehumidification systems". Applied Thermal Engineering. 236: 121514. doi:10.1016/j.applthermaleng.2023.121514. ISSN   1359-4311.
  3. J. H. Lienhard IV; J. H. Lienhard V (August 14, 2020). A Heat Transfer Textbook. Phlogiston Press. pp. 121–127.
  4. Shah, Ramesh; Sekulic, Dusan (2003). Fundamentals of Heat Exchanger Design (1 ed.). Hoboken, NJ: John Wiley & Sons. ISBN   0-471-32171-0.
  5. Fix, Andrew J.; Gupta, Shivam; Braun, James E.; Warsinger, David M. (2023-01-15). "Demonstrating non-isothermal vacuum membrane air dehumidification for efficient next-generation air conditioning". Energy Conversion and Management. 276: 116491. doi: 10.1016/j.enconman.2022.116491 . ISSN   0196-8904.