Which statement defines the heat capacity of a sample?

For an ideal gas component i, the specific heat capacity at constant pressure can be derived as follows,

(35)cpi0=Bi+2CiT+3DiT2+4EiT3+5FiT4

where, Bi, Ci, Di, Ei, Fi are constants related to component i.

For an ideal gas mixture,

(36)cp0=∑iyicpi0

For a real gas, the specific heat capacity at constant volume can be calculated by,

(37)cv=cv0+∫0ρ∂cv∂ρTdρ=cv0+∫0ρ−Tρ2∂2p∂T2ρdρ

Substituting the EOS of natural gas into Eq. (37) can yield the specific heat capacity at constant volume (Table 4). According to the relation between cv and cp (see Eq. (34)), the specific heat capacity at constant pressure can also be easily calculated.

Table 4. Calculation formulas for specific heat capacity of natural gas.

EOSCalculation formulas for specific heat capacityRKcv=cv0+0.5abT0.5ln1+bρSRKcv=cv0+d2adT2Tbln1+bρPRcv=cv0−T22bd2adT2ln1+1+2bρ1+1−2bρBWRScv=cv0+6C0T3−12D0T4+20E0T5ρ+dT2ρ2−2αd5T2ρ5+3cγT3γρ2+2e−γρ2−2

The ratio of the specific heat capacity at constant pressure to that at constant volume is called the heat capacity ratio,

(38)k=cpcv

Read moreNavigate Down

View chapterPurchase book

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780128244951000048

From Thermostatics to Non-equilibrium Thermodynamics

Michel Feidt, in Finite Physical Dimensions Optimal Thermodynamics 1, 2017

1.3.1 Calorimetry and calorimetric coefficients

1.3.1.1 Standard definitions

Up to a second order, the expression of elementary heat exchanged when passing from an equilibrium state to an infinitely close equilibrium state takes one of the following forms:

[1.19]δQ=CVdT+lTdV

[1.20]δQ=CPdT+hTdP

[1.21]δQ=λVdP+μPdV

The six calorimetric coefficients involved in these three equations are characteristic to the thermodynamic system. When considered with respect to unit mass, these same coefficients are called specific coefficients; they intrinsically characterize pure substances. These various coefficients are interrelated.

Since dU, dH and dS are total differentials, they lead to Clapeyron relations:

[1.22]lT=T∂P∂TV

[1.23]hT=−T∂V∂TP

then the generalized Mayer’s relation:

[1.24]CP−CV=T∂P∂TV∂V∂TP

Note

γ=CPCV.

1.3.1.2 General definition

Specific heat capacity is the amount of heat to be supplied to (or taken out of) the unit mass of a system in order to increase (or decrease) its temperature by one degree in a thermodynamic process in which quantity X is imposed, according to:

[1.25]δQ=CXdT

Let us note that the heat involved in the process corresponds to a sensible heat. A second type of heat is the latent heat; it is characteristic to phase changes. Vaporization and condensation latent heats play a very important role in many engines.

Read moreNavigate Down

View chapterPurchase book

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9781785482328500017

Thermodynamics and heat transfer

J. Carvill, in Mechanical Engineer's Data Handbook, 1993

3.4.4 Specific heat capacities

Specific heat capacity of solids and liquids (kJ kg−1 K−1)

Aluminium0.897Oil, machine1.676Aluminium bronze0.897Paraffin2.100Brass0.377Paraffin wax2.140Bronze0.343Petroleum2.140Cadmium0.235Phosphorus0.796Constantan0.410Platinum0.133Copper0.384Rubber2.010Ethanol2.940Salt, common0.880(ethyl alcohol)Sand0.796Glass: crown0.670Seawater3.940flint0.503Silica0.800Pyrex0.753Silicon0.737Gold0.129Silver0.236Graphite0.838Tin0.220Ice2.100Titanium0.523Iron: cast0.420Tungsten0.142pure0.447Turpentine1.760Kerosene2.100Uranium0.116Lead0.130Vanadium0.482Magnesia0.930Water4.196Magnesium1.030Water, heavy4.221Mercury0.138Wood (typical)2.0 toMolybdenum0.2723.0Nickel0.457Zinc0.388

Specific heat capacity of gases, gas constant and molecular weight (at normal pressure and temperature)

GasSpecific heats cp(kJ kg−1 K−1) cvγ=cpcvGas constant, R (kJ kg−1K−1)Molecular weight, MAir1.0050.7181.40.287128.96Ammonia2.1911.6631.320.52815.75Argon0.52340.31361.6680.208140Butane1.681.511.110.1758Carbon dioxide0.84570.65731.290.188944Carbon monoxide1.0410.74491.3980.296828Chlorine0.5110.3831.330.12865Ethane1.76681.49471.180.276530Helium5.2343.15681.6592.0774Hydrogen14.32310.19651.4054.1242Hydrogen chloride0.8130.5831.400.23036.15Methane2.23161.71241.300.518316Nitrogen1.0400.74361.400.296828Nitrous oxide0.9280.7081.310.22037.8Oxygen0.91820.65861.3940.259832Propane1.69151.5071.120.188644Sulphur dioxide0.64480.51501.250.129864

Read moreNavigate Down

View chapterPurchase book

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B978008051135150008X

Polymer Characterization

C. Schick, in Polymer Science: A Comprehensive Reference, 2012

2.31.2.2.1 Linear scan

The most common mode of operation in DSC is heating or cooling at constant rates. The primary outcome of such an experiment is a plot of the heat flow rate versus time. If the temperature of the sample position is known, then data can also be represented as the heat flow rate versus temperature. (One should know that generally a temperature near the sample is measured and not the sample temperature itself.) Figure 2 shows a typical example.

Which statement defines the heat capacity of a sample?

Figure 2. Temperature profile and measured heat flow rate for (a) empty pans, (b) sapphire calibration standard (31.3 mg), and (c) initially amorphous PEEK (29 mg). Heating rate β = 20 K min−1.

Data from PerkinElmer Pyris Diamond DSC. Reproduced with permission from Schick, C. Anal. Bioanal. Chem. 2009, 395, 1589–1611.35

From the heat flow rate curves shown in Figure 2, specific heat capacity cp(T) can be obtained as follows

[8]cp(T)=cp,sapphire(T)msapphireβmsampleβΦsample(T)−Φempty(T)Φsapphire(T)−Φempty(T)=K(T)Φsample(T)−Φempty(T)msampleβ

with

K(T)=cp,sapphire(T)msapphireβΦsapphire(T)−Φempty(T)

where K(T) is a temperature-dependent calibration factor, which can be stored for future use. Here, all measurements are collected at the same scanning rate. The isotherms at the beginning and the end of the scan are used to correct for small changes in heat losses between empty, sapphire, and sample measurements by aligning these parts of the curves. Small changes in losses are unavoidable, because the thermal properties, such as thermal conductivity, of the samples are different. On the other hand, inspection of the heat flow rate at the isotherms allows us to check the correct placement and thermal contacts of all the parts of the measuring system moved during sample changes. Especially the high temperature isotherm should not vary too much between successive measurements.

Specific heat capacity is the most useful quantity available from DSC because it is directly related to sample properties and, according to eqns [1]–[5], directly linked to stability and order. Nevertheless, often only heat flow rate, as obtained from a single sample measurement, is presented. There are several reasons why this should not be presented:

1.

Each heat flow rate graph needs indication of endothermic or exothermic direction because plot direction is not standardized.

2.

Curves measured at different scanning rates are not easy to compare.

3.

If not divided by the sample mass, curves for different samples cannot be compared.

4.

If empty pan measurements are not subtracted, traces may be curved and baseline construction for peak integration may be difficult.

5.

If the heat flow rate calibration factor K(T) is temperature dependent, the obtained heat of fusions and other such parameters may be erroneous.

Performing corrections (3)–(5) yields specific heat capacity as given by eqn [8]. Because most of the DSC software packages include determination of specific heat capacity according to eqn [8], it is strongly recommended to determine specific heat capacity and not to present heat flow rate curves. Even though presenting specific heat capacity data is preferable, there may be reasons not to do so. The normalization of the heat flow rate curve by the scan rate and the sample mass may result in ‘pseudo cp measurements’, which can be used to determine temperature-dependent crystallinity and other quantities as shown in Reference 8. But there is another very strong argument in favor of presenting specific heat capacity rather than ‘pseudo cp’ or heat flow rate. For more than 200 polymers, specific heat capacity data from 0 to 1000 K are available from the ATHAS Data Bank (ATHAS-DB).36 The data can be used for a comparison of measured data in the glassy or liquid state with the recommended values. This allows an easy check of the quality of the measured data, although one should keep in mind that the accuracy of the recommended data bank data is only about 6%. Figure 3 shows specific heat capacity (according to eqn [8]) calculated from the data shown in Figure 2.

Which statement defines the heat capacity of a sample?

Figure 3. Specific heat capacity versus temperature for an initially amorphous PEEK sample. Data from Figure 2. Reference data (straight lines) for the fully amorphous (liquid) and crystalline (solid) PEEK are available from the ATHAS-DB.36

Reproduced with permission from Schick, C. Anal. Bioanal. Chem. 2009, 395, 1589–1611.35

A more detailed discussion of the evaluation of the curves shown in Figure 3 is given in Reference 35.

Besides scan measurements on heating, DSC allows cooling in a wide range of cooling rates. Depending on the instrument and the temperature range of interest, cooling rates up to 750 K min−1 may be reached (HyperDSC™ PerkinElmer, USA).20,37–39 But generally the temperature range for controlled cooling at the highest rates is limited. Measurements performed in a wide range of heating or cooling rates require an optimization of the experimental conditions. The sample mass should scale inversely with the scanning rate. At low rates when the thermal lag is not an issue, the sample mass should be high to have a good signal-to-noise ratio. At high rates when signals are large, the sample mass should be small to minimize the heat flow to the sample, which is proportional to the rate and causes the thermal lag. Problems related to thermal lag, temperature calibration, and reproducibility in fast scanning DSC experiments were intensively studied and adequate recommendations were made.37,40,41 Figure 4 shows cooling curves in the crystallization range of low-density polyethylene (PE). At rates higher than 200 K min−1, controlled cooling down to 100 °C was not possible because of the limited cooling power of the used mechanical intercooler. If higher cooling rates are needed, liquid nitrogen has to be used. For the lower scanning rates shown in Figure 4, the sample mass must be large enough to ensure a good signal-to-noise ratio. For higher rates, the large sample (4 mg) causes some thermal lag, as discussed in the textbooks and References 37, 40, and 42. It is also seen in the broadening of the crystallization peak at 20 K min−1 compared with the 0.4 mg sample at the same cooling rate. Data as shown in Figure 4 provide information about crystallization kinetics and can be analyzed by using different kinetic models.43–48

Which statement defines the heat capacity of a sample?

Figure 4. Cooling curves in the crystallization range of low-density PE. Samples are of mass 4 mg in a 25 mg aluminum pan for cooling rates up to −20 K min−1 and of mass 0.4 mg in 2 mg aluminum foil for higher cooling rates. Heat capacity is plotted downward.

Data from PerkinElmer Pyris 1 DSC. Reproduced with permission from Schick, C. Anal. Bioanal. Chem. 2009, 395, 1589–1611.35

As shown in Figure 4, DSC has a broad dynamic range that can be extended at least by 1 order of magnitude toward lower rates; in this way it covers 3 orders of magnitude. An extension by several orders of magnitude toward higher rates is discussed in Section 2.31.3.2. The possibility of cooling a sample reasonably fast allows us to study the structure formation in far-from-equilibrium situations such as ‘quasi’-isothermal crystallization at deep undercooling.

Read moreNavigate Down

View chapterPurchase book

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B978044453349400056X

Material Interface of Pantograph and Contact Line

Jiqin Wu, in Pantograph and Contact Line System, 2018

5.2.3 Thermal Performance

When selecting materials for strip and contact wire, the thermal performance of the material will also be considered. The pantograph and overhead contact line system works in a strong current environment. Joule heat, arc heat, and friction heat caused by current flowing through the pantograph and overhead contact line contact point will lead to temperature rise of strip and contact wire. Excessive temperature rise will bring negative impact to performance of strip and contact wire material.

Specific heat capacity, thermal expansion, heat conduction, thermal radiation, and thermoelectric force are all aspects of thermal performance.

5.2.3.1 Specific heat capacity

Specific heat capacity is the energy required to increase temperature of material of a certain mass by 1°C, in the unit of J/(kg·K). Table 5.2 lists the specific heat capacity of several materials.

Table 5.2. Specific Heat Capacities of Several Materials (25°C)

MaterialsSpecific Heat Capacities [J/(kg·K)]Aluminum900Copper389Silver235Brass375Carbon710

5.2.3.2 Thermal expansion

Expansion and contraction of a substance is a common phenomenon. The expansion coefficient is a parameter indicating such property. Usually, the expansion coefficient refers to length variation of a material per unit length under temperature variation by 1K, so it is also called the linear expansion coefficient (1/K) to be distinguished from the volume expansion coefficient indicating volume variation of material in unit volume. Table 5.3 lists linear expansion coefficients of some materials.

Table 5.3. Linear Expansion Coefficients of Some Materials (25°C)

MaterialsLinear Expansion Coefficients (×0.000001/°C)Ordinary cast iron9.2–11.8Iron12–12.5Copper18.5Bronze17.5Brass18.5Aluminum alloy23.8

5.2.3.3 Heat conduction

Applying thermal vibration with more energy from outside to a mass point in thermal vibration at certain temperatures will lead to an increase of thermal vibration of adjacent mass points. In such cases, the wave peak with higher thermal vibration moves to low temperature to transfer large thermal vibration introduced at first via the mass point. This phenomenon is heat conduction, namely energy migration generated by temperature difference between adjacent parts of a material. The constant representing the heat conduction ability of material is called heat conductivity or heat conductivity coefficient, in W/(m·K), namely the heat transferred through unit horizontal cross-sectional area in unit time when at vertical temperature gradient of 1°C/m.

Table 5.4 lists the heat conductivity of some materials at normal temperature.

Table 5.4. Heat Conductivities of Several Materials (25°C)

MaterialsHeat Conductivities [×4.2 × 102W/(m·K)]Copper0.927Brass0.26Aluminum0.488Carbon129

Metal is a good conductor of electricity and heat. This is because free electrons exist in metal. Collisions between free electrons become more and more frequent with a rise of temperature, and their movement will become difficult. So, heat conductivity of metal reduces with a rise of temperature. Impurity in metal will obstruct the motion of free electrons and reduce conductivity, so heat conductivity of alloys is significantly lower, equivalent to 15%–70% of the base phase metal.

5.2.3.4 Heat resistance

Heat resistance is an important property in material application. The melting point of a material can reflect the heat resistance of the material. The temperature at fusion welding is called the melting point. Usually, the larger the intermolecular force in a material’s structure, the higher the melting point will be. Table 5.5 lists the melting points of several materials.

Table 5.5. Melting Points of Several Materials

MaterialsMelting Points (°C)Copper1083Steel1515Aluminum660Graphite≈3700

Read moreNavigate Down

View chapterPurchase book

Read full chapter

URL: https://www.sciencedirect.com/science/article/pii/B9780128128862000057

Further Considerations in Field Modeling

In Computational Fluid Dynamics in Fire Engineering, 2009

Specific Heat Capacity of Dry Virgin Wood

The specific heat capacity of dry virgin wood as a function of temperature has been determined by Atreya (1983) for a temperature range from 0°C to 140°C. Fredlund (1988) has assumed that the relationship is valid even for temperatures above 140°C. The expression for the specific heat capacity of dry virgin wood as a linear function of temperature is given by

(4.11.37)Cpw=Cpw,o+Cpw,mTs

where Cpw,o = 1.4 kJ kg−1 K−1 and Cpw,m = 3.0 × 10−4 kJ kg−1 K−2. It is nonetheless noted that a constant specific heat capacity for dry virgin wood independent of temperature has also been assumed by a number of investigators such as Kanury and Blackshear (1970a), Kung (1972), Kung and Kalelkar (1973), Chan et al. (1985), Alves and Figueiredo (1989), Bonnefoy et al. (1993), and Di Blasi (1994a). Values ranging from 1.386 kJ kg−1 K−1 to 2.52 kJ kg−1 K−1 with most of them larger than 2.0 kJ kg−1 K−1 have been typically employed.

What statement defines the heat capacity of sample?

The heat capacity of a substance can be defined as the amount of heat required to change its temperature by one degree.

What defines heat capacity?

heat capacity, ratio of heat absorbed by a material to the temperature change. It is usually expressed as calories per degree in terms of the actual amount of material being considered, most commonly a mole (the molecular weight in grams). The heat capacity in calories per gram is called specific heat.

Which statement defines the heat?

Answer and Explanation: Heat is the amount of energy generated when there is an increase in the temperature of that substance.

What is heat capacity quizlet?

Define heat capacity. It is the amount of heat energy needed to increase the temperature of an object by 1 degree celsius.