Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Solar Energy Conversion: An Introductory Course
Solar Energy Conversion: An Introductory Course
Solar Energy Conversion: An Introductory Course
Ebook1,804 pages

Solar Energy Conversion: An Introductory Course

Rating: 0 out of 5 stars

()

Read preview

About this ebook

Solar Energy Conversion: An Introductory Course is a collection of papers that deals with the technical, mechanical, and operation concerns in converting solar energy. The title first details solar radiation, and then proceeds to discussing solar collectors. Next, the selection covers selective surfaces and the thermal regulation of buildings. The text also talks about planning of solar architectures. The next part tackles topics about the direct conversion of solar energy. Part VII discusses the control and measurement of collected solar energy, while Part VIII covers bioconversion and biomass. The book will be of great use to engineering and science students. Professionals involved in the research and development of solar technology will also benefit from the text.
LanguageEnglish
Release dateSep 3, 2013
ISBN9781483189284
Solar Energy Conversion: An Introductory Course

Related to Solar Energy Conversion

Mechanical Engineering For You

View More

Reviews for Solar Energy Conversion

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Solar Energy Conversion - A. E. Dixon

    CENTRATOM

    INTRODUCTION

    The lectures contained in this volume were delivered at the Fifth Course on Solar Energy Conversion, held at the University of Waterloo August 6–19, 1978. The summer school was attended by 120 delegates from more than 30 countries. Most delegates were either graduate students in one of the many areas of Solar Energy Conversion, or practicing scientists or engineers. The lectures on each topic included review lectures where they were thought to be necessary, so this volume should be useful to senior undergraduates and graduate students in science and engineering, and to engineers and scientists who need a broad overview of the field.

    The Fifth Course received financial support from the University of Waterloo and the United States Department of Energy. Partial support was granted to delegates from underdeveloped countries to help pay their living and travel expenses using grants received from the Canadian International Development Agency and the International Centre for Theoretical Physics (Trieste, Italy). The course was sponsored by the Department of Physics of the University of Waterloo and by the International Centre for Theoretical Physics. The international organizing committee was headed by Prof. A.A.M. Sayigh, of the College of Engineering, University of Riyadh, Saudi Arabia.

    We would like to thank Susan Reidel for her help in re-typing, correcting, and assembling manuscripts, and Lynn Shwadchuck for touching up or re-drawing several diagrams. Cover graphics created by George Roth.

    June, 1979

    A.E. Dixon and J.D. Leslie,     Waterloo

    SOLAR RADIATION

    1

    BASICS OF SOLAR ENERGY

    A.A.M. Sayigh,     College of Engineering, University of Riyadh, Riyadh, Saudi Arabia

    Publisher Summary

    This chapter discusses the basics of solar energy. The sun is part of the Milky Way, which is a spiral composed of over 10¹⁰ stars. The horizontal system is preferred when calculating the position of the sun with respect to the geographical coordinates on the earth. In this system, the reference plane is the horizon of the observer, that is, the plane passing through the observer and normal to the vertical. The vertical circles mean any of the great (maximum) circles of the sphere passing through the zenith of the observer. A particular vertical circle is the meridian—the vertical circle passing also through the celestial poles. Hour circle is the great circle normal to the celestial equator and passing through the sun; it is also called the declination circle as the angular distance from the celestial equator to the sun measured on this circle corresponds to the declination. This distance is constant during the day because the apparent path of the sun is always a circle parallel to the celestial equator.

    1 INTRODUCTION

    The sun is a grain of sand in a whirling desert storm. It is part of the Milky Way which is a spiral composed of over 10¹⁰ stars. The basic characteristics of the sun are:

    A more detailed description of the various zones of the sun is given in reference (1).

    2 COMMON TERMINOLOGIES

    The horizontal system (Figures 1 and 2) is preferred when calculating the position of the sun with respect to the geographical coordinates on the earth. In this system, the reference plane is the horizon of the observer, i.e. the plane passing through the observer and normal to the vertical. The elements in this system are:

    FIG. 1 The celestial vault represented in the horizontal system.

    FIG. 2 The astronomical spherical triangle NcZS, relative to the earth’s coordinates. Point S, the interception of the vertical circle with the hour circle, is not labeled for graphical reasons.

    : The point at which the sphere intercepts the upward vertical axis passing through the observer.

    Nadir, N: The point of the celestial sphere diametrically opposite to the zenith.

    Celestial poles, NC and SC: The zeniths of the terrestrial poles.

    Vertical circles: Any of the great (maximum) circles of the sphere passing through the zenith of the observer. A particular vertical circle is the meridian: the vertical circle passing also through the celestial poles.

    Celestial equator: The great circle normal to the earth’s axis.

    Hour circle: The great circle normal to the celestial equator and passing through the sun; it is also called the declination circle as the angular distance from the celestial equator to the sun measured on this circle corresponds to the declination. This distance is constant during the day because the apparent path of the sun is always a circle parallel to the celestial equator.

    Almucantar: The altitude parallels, i.e. every circle parallel to the horizon.

    Altitude, β: The almucantar, i.e. the angular elevation above the horizontal positive towards the zenith, negative towards the nadir.

    Azimuth, A: The bearing, i.e. the angular distance from the meridian to the great circle passing through the zenith of the observer and the celestial body. It is measured on the horizontal from north towards east, from 0° to 360°, positively, or negatively in the opposite direction.

    3 THE SOLAR DECLINATION

    The solar altitude angle measured at noon will differ from the corresponding equinocial angle by an angle of up to ± 23° 17’. This angle is called the solar declination. It is defined as the angular distance from the zenith of the observer at the equator and the sun at solar noon. It is positive when it is north and negative when it is south. The declination reaches its maximum value, +23° 17’, on 21 June (the summer solstice in the northern hemisphere, the winter solstice in the southern hemisphere). The minimum value, −23° 27’, is reached on 20 December. The declination, in degrees, for any given day may be calculated in first approximation with the equation:

    where d represents the number of days passed after the spring equinox (spring referred to the northern hemisphere), which is 21 March (2). Figure 3 shows the declination angle versus days after the equinox. Table I shows the solar declination for zero hour Greenwich mean time.

    TABLE I

    THE SOLAR DECLINATION*

    *US Naval Observatory, The American Ephemeris and Nautical Almanac for the Year 1950, Washington, 1948. All the above data are for oh Greenwich civil time in the year 1950; the variations of these data from year to year are negligible for solar engineering purposes; the largest variation occurs through the four-year leap-year cycle.

    FIG. 3 The declination, calculated and tabulated.

    4 THE LENGTH OF THE DAY (Z)

    The length of the day can be calculated from the equation:

    where δ is the declination angle of the place, and ϕ is the latitude of the place. Figure 4 shows a nomogram for calculating the length of the day (Z) in hours (3).

    FIG. 4 A NOMOGRAM FOR CALCULATING THE LENGTH OF THE DAY (Z) IN HOURS.

    5. REFERENCES

    1. Sayigh, A.A.M., eds. Solar Energy Engineering. Academic Press, London, 1977.

    2. A.A.M. Sayigh, Formulae for total and diffuse radiation, 3rd Course in Solar Energy Conversion, Sponsored by International College of Applied Physics, Catania, Italy, 5–17 September 1976.

    3. Whillier, A., Solar radiation graphs. Solar Energy 1965; 9:165

    2

    CHARACTERISTICS OF SOLAR RADIATION

    A.A.M. Sayigh,     College of Engineering, University of Riyadh, Riyadh, Saudi Arabia

    Publisher Summary

    This chapter discusses the characteristics of solar radiation. The most important parameter in the survey of solar radiation measurement is the energy received outside the earth’s atmosphere. This parameter has a degree of constancy when compared to that received on the ground. Therefore, it is called the solar constant. It is defined as the energy received from the sun on unit area exposed normally to the sun’s rays at the average sun–earth distance in the absence of the earth’s atmosphere. The 1970 ISES Congress in Melbourne accepted the value of 1353 W/m² for the solar constant with an error of about ± 1.5%. The variability of solar energy incident on a collector surface on the ground is considerably greater than that of the extraterrestrial solar energy. On a clear sunshine day, the energy increases from zero at sunrise to a maximum at solar noon and decreases to zero at sunset. At any moment, clouds may intercept the sun and decrease the energy to a low value because of the diffuse radiation.

    1 INTRODUCTION

    The earth receives solar energy at a range of 5.4 × 10²⁴ J/year. This is equivalent to about 30,000 times the sources of energy used at the present time. Harnessing this power requires knowledge of the nature of solar insolation, the factors which influence its intensity and tools which utilise such energy. For example, in building applications, the total solar radiation (global radiation) is the source of energy. The intensity of this source depends on whether the radiation is direct or diffuse, or both. As for the tools which utilise the sun’s energy, the most efficient are flat plate collectors. In the case of solar concentrators, direct solar radiation is the source.

    2 THE SOLAR CONSTANT AND EXTRATERRESTRIAL SOLAR SPECTRUM

    The most important parameter in the survey of solar radiation measurement is the energy received outside the earth’s atmosphere. This parameter has a degree of constancy, when compared to that received on the ground. Therefore, it is called the solar constant. It is defined as the energy received from the sun on unit area exposed normally to the sun’s rays at the average sun-earth distance in the absence of the earth’s atmosphere. The 1970 ISES Congress in Melbourne accepted the value of 1353 W/m² for the solar constant, with an error of about ± 1.5%. This value is an average value for nine long series of measurements, all made from high altitude platforms (Convair 990, balloons, X-15 aircraft, and Mariner Mars probe) during the period 1967–1970. Figure 1 shows these measurements (1). A more recent suggestion states the value of solar constant to be between 1368 and 1377 W/m² as deduced from Table I. The extraterrestrial solar spectrum is shown in Figure 2. The spectrum consists of ultra-violet region from 0.115 to 0.405 micron, the percentage energy collected in this region being 9.293% (2), visible region from 0.405 to 0.740 micron, the percentage of energy collected in this region being 41.476%, and infrared region from 0.740 to 1000 micron, the percentage of energy collected in this region being 49.231%. It is worth mentioning here that the percentage of energy in the infrared region, from 0.740 to 5.000 microns, is 48.743%. Table II shows these various percentages with their respective wavelengths.

    TABLE I

    SOLAR CONSTANT RECENT VALUES

    TABLE II

    SOLAR SPECTRAL IRRADIANCE AT VARIOUS WAVELENGTHS BASED ON SOLAR CONSTANT OF 1353 Wm−2

    Figure 1 Values of the Solar Constant Derived from High Altitude Measurements

    Figure 2 Solar Spectrum Irradiance, Standard curve, Solar Constant 1353 W m−2

    3 TOTAL AND SPECTRAL SOLAR IRRADIANCE AT GROUND LEVEL

    The variability of solar energy incident on a collector surface on the ground is considerably greater than that of the extraterrestrial solar energy. On a clear sunshine day, the energy increases from zero at sunrise to a maximum at solar noon and decreases to zero at sunset. At any moment, clouds may intercept the sun and decrease the energy to a low value due to the diffuse radiation. Figures 3 and 4 show such variations for the 13 and 14 May 1971. The two figures are based on measurements made on two successive days as part of the GSFC international comparison of working standard pyranometers. The instrument was an Eppley pyranometer, model 2, mounted on a roof top. The readings were taken every four seconds. On 13 May, it was overcast during most of the morning. A short interval of sunshine in the afternoon was followed by a heavy cloudburst at 18.00 hours. The next day was one of the relatively clear sunshine with a few passing clouds. The high value of solar irradiance was near 1200 W/m², after 14.00 hours on 13 May, is 30% higher than a clear day for the given solar elevation; this is obviously due to reflection from the clouds. Such abnormally high values are of short duration.

    Figure 3 Global Irradiance Due to the Sun and Sky on a Horizontal Surface, Measured at GSFC on May 13, 1971. Total Energy Received During the Day, 175 cal cm−2(732 Joules cm−2)

    Figure 4 Global Irradiance Due to the Sun and Sky on a Horizontal Surface, Measured at GSFC on May 14, 1971.

    As for attenuation of solar energy, this depends on Rayleigh attenuation coefficient, C1, and the ozone attenuation coefficient, C2. Both factors are based on the data developed by reference (3). They are valid for US standard atmosphere (2). For atmospheric turbidity, Angstrom has developed an equation:

    (1)

    where β is the Angstrom coefficient;

    α is the wavelength exponent; and

    λ is the wavelength in μm.

    This equation permits a greater flexibility in choosing α and β parameters corresponding to different levels of atmospheric pollution. In the infrared, a fourth parameter to account for the molecular absorption bands is required (4).

    It is important to define a common term, AIR MASS, now: it is the ratio of the path length of the radiation through the atmosphere at any given angle to the sea level path straight through the atmosphere (vertically). Air mass zero refers to the absence of atmospheric attenuation at one astronomical unit from the sun.

    For simplicity, it can be assumed that the zenith angle is small and, therefore, the air mass (m) can be expressed as:

    (2)

    where Z is the zenith angle. For large value of Z, account has to be taken of the curvature of the solar ray due to refraction in increasingly denser atmospheric layers; Figure 5 shows the effect of various air mass on solar irradiance for US standard atmospheric turbidity. That is, H2O = 20 mm, O3 = 3.4 mm, α = 1.3, β = 0.02, and the various air masses were: air mass zero which is for no atmosphere, air mass one which is corresponding to Z = 0, air mass four corresponding to Z = 75.5°, air mass seven corresponding to Z = 81.7°, and air mass ten corresponding to Z = 84.3°. Figure 6 shows four different curves related to solar spectrum irradiance.

    FIG. 5 SOLAR IRRADIANCE FOR DIFFERENT AIR MASS VALUES U.S. STD. ATMOSPHERE; H2O = 20mm; O3 = 3.4mm (α = 1.3; β 0.02)

    FIG. 6 Four curves related to solar spectral irradiance.

    The investigation of the solar radiation on inclined surfaces and their predictions from the horizontal component of the radiation which is well established in most parts of the world, is of great importance in utilising solar radiation effectively in such places.

    4 SOLAR RADIATION ON INCLINED SURFACE

    The problem which deals with direct solar radiation is well explained by various authors (5), (6), (7), (8). Using Figure 7, the solar radiation flux, Hi, falling on an arbitrarily oriented inclined surface, the inclination angle of the surface is b, can be expressed by the solar radiation component normal to the surface Hn using the following expression (9):

    Figure 7 Angle Diagram for finding Solar Radiation or, Sloped Surfaces

    (3)

    and:

    (4)

    = Z p, where p is the azimuth of the plane, and Z is the sun azimuth from the south. α and Z can be determined from the following equations:

    (5)

    (6)

    (7)

    where L is the northern latitude of the place, δ is the sun’s declination, and h is the hour angle (zero at noon time, positive in the afternoon and negative before noon). Therefore, using the previous equations, equation (2) can be written as:

    (8)

    For a plane on any north-south slope, in other words truly facing south:

    (9)

    and is h = 0, which is at noon, then:

    (10)

    For example, in the town of Riyadh, Saudi Arabia, L − 25°N, and for a collector placed horizontally at noon during the middle of the months:

    (11)

    Table III shows the values of δ and the various components of solar radiation at various angles. The angles were zero, normal, 15°, 30°, 45°, 60° and 90°, respectively. Figures 8 and 9 show these results in graphical form (9). Figure 10 shows a monogram for calculating the length of the day which includes a curve for the declination angle (10).

    TABLE III

    DECLINATION ANGLE AND VARIOUS SOLAR RADIATION VALUES FALLING ON DIFFERENT SLOPED SURFACES IN RIYADH (gm.cal/cm²/day)

    FIG 8 SOLAR RADIATION FALLING ON VARIOUS SLOPED SURFACES IN RIYADH

    FIG. 9 AVERAGE DAILY SOLAR RADIATION ON SLOPED SURFACES IN RIYADH

    Fig 10 A NOMOGRAM FOR CALCULATING THE LENGTH OF THE DAY (Z) IN HOURS.

    Equation (7) is the general expression of the dependence of incoming solar radiation falling on an inclined plane with orientation determined by angles b and p for any latitude L and at various moments of the day (the hourly angle h), or at any time during the year (the sun’s declination δ).

    If the plane is horizontal, then b = 0 and Hn is the direct radiation on a horizontal plane, then equation (7) becomes:

    (12)

    Also:

    (13)

    is the relative azimuth between sun and normal to the receiving surface.

    If equation (12) is rewritten as:

    (14)

    the characteristic line is obtained by plotting Ψ(H) for given values of α and b . Figure 11 shows this plot. It is a straight line of slope unity and passing through the points (1,1), (0,0) and (−1,−1). This is only true if b is always less than α.

    Fig. 11 Ideal Radiation Line for Direct Radiation only

    The scale for cos is in the reverse direction to that normally used. This is justified because, if one starts with the plate facing the sun, and progressive relative motion being assumed until the sun is behind the plate.

    If b is greater than α, then the sun rays cease to strike the front surface of the plate, for example if b = 90° or cos = 0, irrespective of α.

    For values of b less than 90°, but greater than α, the ratio Hi/HO becomes zero. This is when the sun’s rays pass tangentially to the plane when:

    5 DIFFUSE RADIATION

    For clear sky, this component of solar radiation depends mostly on the air mass (m) and atmospheric turbidity, water vapour, dust contents and aerosols. This component is equivalent to 16% of the total radiation, as a rough estimate falling on a horizontal surface at noon. It rises to 25% five hours later in the afternoon. These values are for mid-summer in a temperate climate (7). If the distribution of diffuse radiation is uniform over the whole of the visible sky hemisphere, then the diffuse radiation falling on an inclined plane is:

    (15)

    For further work regarding diffuse radiation on inclined surfaces, see K. Ya. Kondratyev (11). Figure 12 shows the ratio of diffuse radiation on inclined surface to that on horizontal surface for solar altitude of 47°, at Woolwich, UK.

    Fig. 12 Ratio of Diffuse Radiaton on Inclined and Horizontal Surfaces for Woolwich, June 26, 1964 - from Heywood (1965) - for Solar Altitude of 47°

    6 REFLECTED RADIATION

    Reflected radiation, or re-radiation from the ground in most cases, is neglected, especially on a surface with inclination b less than 60°. Heywood thought this factor could be as large as 8% in a vertical surface. Care must be taken in using radiation equipment so that no reflected radiation from adjacent walls or ground is present. However, it is difficult to separate this component from diffuse radiation component. An extensive analysis of diffuse radiation, as well as reflected radiation, is given by reference (12). The diffuse radiation on an inclined surface and its intensity can be expressed as:

    (16)

    where J1) is the intensity of radiation scattered in the direction determined by coordinates α1 relative azimuth direction α1(Zrelative to the plane of the horizon. In the case of isotropic scattering radiation, J) = J = constant; therefore, equation (15) will be:

    (17)

    Another relationship can be written for reflected radiation:

    (18)

    1) is the reflected radiation intensity:

    ) = Jref = constant), then equation (18) becomes:

    (19)

    7 TOTAL RADIATION OR GLOBAL RADIATION

    It is of great importance in solar energy technology to know the total radiation falling on an inclined plane. In calculating the total fluxes, the non-isotropy of scattered radiation should be taken into account. It is apparent that, when the sky is clear and the solar altitude is small, a plane facing the sun will receive a substantial amount of scattered radiation compared with total radiation. For large solar altitude, the relative contribution of scattered radiation is insignificant. The global radiation flux to slopes may be calculated under the assumption of the isotropy of scattered radiation. The experimental values of total radiation, Ei and EO on cloudless days, showed a straight line, but this is different from the ideal in having a slope less than unity and not passing through the point (0,0). Numerous experimental readings led to the following relationship:

    (20)

    where C = Ψ(E) when cos = 0, and M is the slope of the characteristic line. Figure 13 shows typical characteristic lines for two angles of inclination of plane and three altitudes. Estimation of total solar radiation on a horizontal surface has been carried out by reference (10).

    Fig. 13 Radiation lines for total radiation on surfaces at 90 and 40 deg to the horizontal and for various solar altitudes

    8 THE EFFECT OF TURBIDITY ON SOLAR INTENSITY

    Turbidity (τα) effect is almost inversely proportional to the total solar radiation reaching the ground. This is clearly shown in Figures 14 and 15. Unsworth, reference (13), found that, from a long series of measurement of τα in Britain, τα depends on the air stream prevailing over the area. It was also stated that τα = 0.05 in clear polar air streams, while τα = 0.35 in air stream of continental origin and polluted urban areas. τα depends mostly on the amount of aerosol in the atmosphere.

    Figure 14 Calculated direct irradiance at normal incidence on June 22 at latitude 52°N for turbidities τa = 0.1, and τa = 0.4

    9 THE SURFACE ALBEDO, αg

    One of the important parameters in dealing with solar radiation intensity is the reflectance of various surfaces. This parameter is called surface albedo (αg). It is not exactly uniform for a given surface, but for practical purposes is assumed uniform. It is also a function of the air mass or solar zenith angle Z. Table IV shows the albedo for various surfaces at various solar elevations (14).

    TABLE IV

    THE ALBEDO OF VARIOUS SURFACES AT VARIOUS SOLAR ELEVATIONS

    10 ESTIMATION OF TOTAL SOLAR RADIATION

    10.1 The Use of Sunshine Hours

    Masson (1966) plotted the total solar radiation, H, on a horizontal plane in langleys (1y) per day against the hours of sunshine per day, S, and fitted a hyperbolic segment to the result. The segment tends to be a straight line if the hours of sunshine are equal or greater than 7 hours. The results were expressed in mathematical form as:

    or:

    10.2 The Use of the Ratio of Sunshine Hours to Length of Day

    One of the earliest expressions was Angstrom’s regression (1924), which is:

    where A’ and B’ are arbitrary constants (Fritz (1951) suggested that A’ = 0.35 and B’ = 0.61, HO is the monthly average horizontal solar radiation in W/m−2, where δ is the declination angle of the place, and ϕ the latitude of the place). An easier way of calculating Z is from a nomogram that was developed by Whillier (1965) and is shown in Figure 10. A better form of the previous equation is suggested by reference (2):

    where A and B are arbitrary constants, and HO is the average monthly insolation at the top of the atmosphere.

    10.3 The Use of Angström’s Equation with an Extra One or Two Parameters

    Bennett (2) derives two sets of equations. One set for Canadian stations for June and December are, respectively:

    and:

    where HO equals 1.98 1g/min, and S, is the monthly mean daily percentage of possible sunshine.

    The second set is for North America in which he introduces the station elevation, h, in feet, i.e. for June and December, are, respectively:

    and:

    In another article, Bennett (2) writes these equations as:

    and for the United States, A varies from 188 in April to 291 in August, B varies from 3.768 in August to 5.574 in April, and C varies from 0.00130 in July to 0.0226 in December. Swartman & Ogunlade (2) use the relative humidity, R, in three different equations:

    and:

    11 TOTAL RADIATION - NEWLY PROPOSED FORMULA

    where:

    ϕ in radians, which is the latitude.

    where:

    δ is the declination angle in degrees;and:

    d is the number of days after the spring equinox (which is 21 March for the northern hemisphere, and 21 September for the southern hemisphere).

    Ψi,j is the relative humidity factor in graphical form (see reference (10)).

    Twelve different locations all over the world, from latitude −40° to +60°, were tested in using this formula, and the results were excellent. For further details, see reference (10).

    12 CONCLUSION

    It is essential to know how much solar energy is available at a given location before utilising it. In buildings, the most effective absorber is the flat plate collector. This is due to its ability of utilising diffuse radiation as well as direct radiation. The correct orientation of the solar absorber is also essential in obtaining maximum solar radiation intensities. Obviously, the sun tracking absorber is best, but due to the expense of such a device and to the impractability in most buildings, a fixed orientation for the collectors in a building is preferred.

    13. REFERENCES

    1. M.P. Thekaekara, Survey of quantitative data on the solar energy and its spectral distribution, COMPLES Conference, Dhahran, (1975).

    2. Sayigh, A.A.M. Solar Energy Engineering, 1st edition. Academic Press, 1977.

    3. Elterman, L.UV, visible and IR attenuation for altitude to 50 km. Office of Aerospace Research, US Air Force, 1968. [AFCR-68–0153].

    4. Gates, D.M., Harrop, W.J. Infrared transmission of the atmospheric to solar radiation. App. Optics. 1963; 2:887.

    5. Kondratyev, K.Ya.Radiation in the Atmosphere. New York: Academic Press, 1969.

    6. A.M. Zarem, and D.D. Erway, Introduction to the utilisation of solar energy, University of California Engineering & Science Extension Series, (1963).

    7. Heywood, H. The computation of solar radiation intensities – Part I. Solar Energy. 9(4), 1965.

    8. Heywood, H. The computation of solar radiation intensities - Part II. Solar Energy. 10(1), 1966.

    9. A.A.M. Sayigh, The technology of flat plate collectors, 4th Course on Solar Energy Conversion, International Centre for Theoretical Physics, Trieste, 6–24 September 1977.

    10. A.A.M. Sayigh, Estimation of total radiation intensities – A universal formula, LAGA/AMAP Joint Assembly Conference, Seattle, USA, 22 August – 3 September 1977.

    11. K.Ya. Kondratyev, and M.P. Manolova, Scattered and global radiation income to inclined surfaces in the presence of snow cover, Vestnik of Leningrad State University, (16), 67, (1960).

    12. K.Ya. Kondratyev, and M.P. Fedorova, Radiation regime of inclined surfaces, UNESCO/World Meteorological Organisation Solar Energy Symposium, Geneva, Switzerland, 30 August – 3 September 1976.

    13. Unsworth, M.H., Monteith, J.L. Aerosol and solar radiation in Britain. Quart. J. R. Met. Soc. 1972; 98:778.

    14. Paltridge, G.W., Platt, C.M.R. Radiative Processes in Meteorology and Climatology. 1st edition, Elsevier Scientific Publishing Company; 1976.

    3

    METHODS FOR THE ESTIMATION OF SOLAR ENERGY ON VERTICAL AND INCLINED SURFACES

    J.K. Page,     Department of Building Sciences, University of Sheffield, Sheffield, S102TN, U.K.

    SUMMARY

    This Chapter presents two methods for the assessment of monthly mean solar radiation falling on vertical and inclined surfaces anywhere in the inhabited world. Both methods involve splitting the incident flux into three components, direct, sky diffuse and ground reflected diffuse. Both methods also use mean monthly sunshine data as their starting point. The first method is relatively simple and provides daily values of the irradiation on slopes. The second method involves more complex mathematical modelling and is dependent on the availability of digital computing facilities. It rapidly produces hourly values of the mean monthly irradiance on slopes which check extremely well with field observations. This model does not assume isotropic radiation from the sky. The programs described form part of a set of climatological programs being developed in the University of Sheffield for the systematic assessment of energy gains and losses from buildings under different types of weather. The second part of this lecture is based on the work of a team consisting of Dr. G.G. Rodgers, Mr. C.G. Souster and Miss Jayne Thompson who work in a team led by the author.

    1 INTRODUCTION

    In applied solar energy studies it is useful to have simple methods for estimating the monthly mean daily amounts of solar radiation falling on surfaces of different orientation. While it is quite easy to estimate the amount of radiation received on clear days for any climate whose turbidity is known, simple methods for determining the amount of radiation received averaged over all days have not been so widely available. Most temperate climates have a great deal of cloud as do many equatorial regions and clear days are the exception. Misleading decisions on solar energy applications result if clear day radiation data only are used to select collector orientations. This Chapter starts by briefly summarising a method for the estimation of mean monthly irradiation on inclined surfaces first put forward by the author to the United Nations Conference on New Sources of Energy in Rome in 1961 (1). The method described provides a simple quick technique for estimating mean monthly values of solar irradiation on vertical and inclined surfaces starting from monthly mean values of bright sunshine, or from monthly mean values of global radiation on a horizontal surface. A full discussion of the implications of the non isotropic nature of sky radiance in the prediction of the irradiation on inclined and vertical surfaces follows. The Chapter then goes on to consider more complex computer based methods for the estimation of monthly mean hourly values of the irradiance on slopes developed in Sheffield over the last three years.

    2 A SIMPLE METHOD FOR ESTIMATING THE IRRADIATION ON SLOPES FROM SUNSHINE

    General methodology

    bh is given, supported by appropriate tables covering latitudes 60°N − 40°S. The next step in the prediction process involves estimating the diffuse irradiation from the sky on the surfaces under study. Finally, the ground reflected irradiation has to be added, making appropriate assumptions about the ground albedo. In applied solar energy applications it is, of course, possible to manipulate the ground albedo to significantly increase vertical and inclined surface irradiation.

    The total mean monthly daily irradiation on any inclined surface is thus estimated as a sum of three mean monthly values, the monthly mean direct irradiation from the sun, the monthly mean diffuse irradiation from the sky, and the monthly mean diffuse irradiation reflected from the ground.

    Once such basic calculation procedures have been established it becomes possible to use them to optimise collector orientation for different solar energy applications.

    3 ESTIMATING MONTHLY MEAN DAILY GLOBAL IRRADIATION FROM SUNSHINE DATA

    The relationship between global irradiation and percentage possible sunshine can be estimated in the usually accepted way using regression equations of the Angstrom type:

    (1)

    0 are climatically determined regression constants is often called the percentage possible sunshine.

    ho are given in Table I. These are based on a solar constant of 1353 W/m² at mean solar distance. Values of No are given in Table II.

    TABLE I

    MONTHLY MEAN VALUES FOR CALENDAR MONTHS OF TOTAL DAILY RADIATION ON A HORIZONTAL PLANE ho OUTSIDE THE EARTH’S ATMOSPHERE, LATITUDES 60°N − 40°S. (ASSUMED SOLAR CONSTANT 1353 Wm−2) UNITS: MJm−2 DAY−1

    Note: These values were obtained by planimetry over the monthly periods indicated. Care should be taken in interpolating between latitutdes 40° N and S as the functions go through a maximum in between tabulated values. The table includes the appropriate adjustments for variations in the sun’s distance from the earth.

    TABLE II

    ASTRONOMICAL DAYLENGTH FOR LATITUDES 0 − 60 - UNITS HOURS AND DECIMALS OF HOURS

    Calculated using Dept. of Building Science program SUN1 assuming a correction of 30′ for atmospheric refraction & 20′ for solar disc size making 50 correction in all as in Smithsonian Tables

    Typical values of a and b for a number of stations from different parts of the world are given in o. This affects the values of a and b found by regression analysis.

    TABLE III

    TYPICAL VALUES OF a & b IN REGRESSION EQUATION ) Source: Page (1)

    For sources of data refer to reference 1.

    TABLE IV

    VALUES OF a AND b IN REGRESSION EQUATION FOR UK STATIONS AND IRISH STATIONS REPORTED BY VARIOUS AUTHORS

    There are certain difficulties in this simplified approach which have been discussed in detail in the author’s original paper (o are sometimes used. In this paper values of a and b in the regression equations have been derived using the full daylength from sunrise to sunset and no allowance has been made for the fact that the card on a Campbell Stokes sunshine recorder does not normally burn until the direct intensity reaches 200 W/m².

    4 ESTIMATING MONTHLY MEAN DAILY DIFFUSE IRRADIATION ON A HORIZONTAL SURFACE

    h into these two components in order to estimate the irradiation on vertical and inclined surfaces.

    The basic problem is how, in the absence of local measurements, to separate the diffuse irradiation from the direct irradiation. The author in 1956 found it was fortunately possible to set up reasonably reliable regression equations of the form:

    (2)

    ho = mean monthly daily irradiation on a horizontal plane in the absence of any atmosphere c and d are climatically determined regression constants.

    Typical values of c and d are given in Tables V and VI.

    TABLE V

    REGRESSION EQUATIONS FOR dh/ h AGAINST h/ oh FOR STATIONS BASED ON PERIOD 1965–70, EXCEPT BELFAST, 1969–70, USING MONTHLY MEAN DAILY VALUES OF h AND dh OVER PERIOD, AND DIFFUSE MULTIPLIERS FROM KEW

    *It has just emerged that no corrections were applied for the shading ring correction by Liu and Jordan. This data can no longer be considered valuable.

    TABLE VI

    VALUES OF CONSTANTS c & d IN REGRESSION EQUATION dh/ h = c + d ( h/ oh) FOR STATIONS IN THE AFRICAN CONTINENT

    The mean regression equation for the ten stations scattered across the world, earlier studied by the author (1), was found to be:

    (3)

    5 ESTIMATION OF MEAN MONTHLY DIRECT IRRADIATION ON INCLINED SURFACES

    bh. Using bh, can be considered as an integral of the following form which takes account of the cosine of the angle of incidence i on a horizontal surface which is equal to (90 - α) at time t, where α is the solar altitude.

    where Gbd is the direct irradiance normal to the sun’s rays at time t hours α is the solar altitude at time t hours after start of month. m’ is the number of days in the month.

    Thus

    (4)

    bs = ∫Gbn cos i dt and the clear day daily direct irradiation on the horizontal surface ∫Gbn sin α dt. This ratio is a function of latitude and time of year:

    (5)

    bh as determined from equation 5.

    In the original paper the following surfaces were selected for detailed tabulations for simplified routine calculations using the procedures described above:

    i) vertical surface facing equator,

    ii) vertical surface facing away from equator,

    iii) vertical surface facing east or west,

    iv) 45° slope facing equator,

    v) surface inclined at tilt equal to the latitude.

    The tilt equal to the latitude was included because it is a common tilt recommended by many authors for maximum collection efficiency for solar energy collectors. The standard radiation curve used in these calculations is given in Table VII.

    TABLE VII

    VALUES OF THE DIRECT IRRADIANCE OF THE SOLAR BEAM Gbn ON A SURFACE NORMAL TO THE SUN’S RAYS WITH ALTITUDE OF THE SUN USED TO CONSTRUCT TABLE VII

    Units: Watts m² International Pyrheliometric Scale

    The derived values of Ks have been tabulated at 10° intervals of latitude for 14 dates in the year to simplify practical calculations. Calculated daily values of K obtained by planimetry for latitudes 0 − 60 are given in Table VIII. Ks factors for surfaces other than those actually given in TableVIII may be often derived by simple trigonometrical resolution. The essential condition for deciding whether it is possible to calculate the factor directly from Table VIII or not is to decide whether the direct radiation will reach the surface under consideration throughout the day at that time of year or not. If the radiation is on the surface all day, then simple trigonometrical resolution can be used. However, radiation falling behind a surface must be set as zero otherwise the trigonometrical calculation of instantaneous values will produce negative values for part of the day which, if undetected, in the integrated values, will be subtracted from the positive gains, i.e. energy on the opposite side of the collector will be subtracted from the daily total. For example, an unobstructed inclined surface facing the equator, in the winter half of the year will be insolated throughout the day between the equinoxes. Therefore normal trigonometrical resolution using daily integrated values can be used to estimate Ks for other slopes facing south, for the sun on unobstructed sites will always reach the surface and can never be behind the collector. This statement is not true for the same surface in summer because the sun will move to the poleward side of the inclined surface for part of the day. For example, using Table VIII, the Ks factor for a south facing inclined surface located at 40°N tilted at 60° to the horizontal is equal to (1 × Cos 60 +.93 × Sin 60) on March 15. However, on June 15 the factor is NOT precisely equal to (1 × Cos 60 +.19 × Sin 60) because the integrated values in Table VI cannot take account of the time integration limits and so it is impossible to estimate this correction accurately from Table VI.

    TABLE VIII

    RATIO OF DIRECT RADIATION ON INCLINED SURFACES TO DIRECT RADIATION ON A HORIZONTAL SURFACE AT DIFFERENT TIMES OF THE YEAR FOR LATITUDES 0°–60°, BASED ON STANDARD RADIATION CURVE GIVEN IN TABLE VII

    Column 1. Vert. surface facing Equator: Hor. surface.

    Column 2. Vert. surface facing Pole: Hor. surface.

    Column 3. Vert. surface facing East or West: Hor. surface.

    Column 4. Vert. surface facing Equator tilted at 45° to Hor.: Hor.

    Column 5. Vert. surface facing Equator tilted at an angle to Hor. equal to latitude: Hor. surface.

    NOTE: This table is based on graphical interpolations and the order of accuracy is about - ±2%.

    TABLE VIII

    RATIO OF DIRECT RADIATION ON INCLINED SURFACES TO DIRECT RADIATION ON A HORIZONTAL SURFACE AT DIFFERENT TIMES OF THE YEAR FOR LATITUDES 0°-60°, BASED ON STANDARD RADIATION CURVE GIVEN IN TABLE VII

    Column 1. Vert. surface facing Equator: Hor. surface.

    Column 2. Vert. surface facing Pole: Hor. surface.

    Column 3. Vert. surface facing East or West: Hor. surface.

    Column 4. Vert. surface facing Equator tilted at 45° to Hor.: Hor. surface.

    Column 5. Vert. surface facing Equator tilted at an angle to Hor. equal to latitude: Hor. surface.

    NOTE: This table is based on graphical interpolations and the order of accuracy is about - ±2%.

    6 ESTIMATION OF MEAN MONTHLY DIFFUSE IRRADIATION FROM SKY AND GROUND–ISOTROPIC APPROXIMATIONS

    s on a plane s of slope β will be given by the formulae:

    (6)

    g is the monthly mean diffuse flux reflected from the ground and β is the angle of inclination of the slope to the horizontal plane.

    7 SHORT WAVE RADIATION REFLECTED FROM THE GROUND

    h is the monthly mean daily total irradiation falling on the ground. The ground will normally be insolated for much of the day for the majority of solar energy collectors because these are normally orientated to face towards the Equator to give maximum collection efficiency. However, the precise insolation conditions of the ground need examination for each particular case, especially if overshadowing is expected. Then the ground reflected radiation will be much less because the energy reaching the adjacent ground will be reduced.

    h. Typical values of Y for grass covered surfaces are around 0.25. With snow the value may be very much higher. Suitable values of Y may be selected from the Smithsonian Meteorological Tables (2) and from other sources (3). Table IX contains some typical values of ground albedo useful for making estimates of reflected ground radiation from natural surfaces. In solar energy applications special reflecting surfaces may be used close to structures to augment irradiation. The precise effect will depend on the solid angle subtended by such surfaces as well as their albedo.

    TABLE IX

    ALBEDO OF TYPICAL SURFACES

    *Monteith found a variation in the reflectance of short grass of about 0.22 at a solar altitude of 60° to 0.28 at a solar altitude of 20°. At 10° Roach found values as high as 0.3 to 0.4.

    **Climatic factors, pollution, etc., produce enormous variations. Local assessments of the albedo are desirable.

    The albedo of the ground surface should, if at all possible, be determined locally in view of the wide variations which occur with different types of ground, covered or not as the case may be, with different vegetation with varying reflection characteristics which will change according to locality, season and rainfall. In making measurements of ground albedo it is essential that the total albedo should be measured by using suitable instruments which record over the whole short wave spectrum. The correlation between the visual albedo and the total albedo is often poor, especially over vegetation, because most plants have a relatively high reflectance in the infra red region and often a relatively low one in the visible. Typical reflectance values by Kondratyev, Mironova and Daeva (4) show values for grass of the order of 10–15% in the visible region and 40–50% in the near infra red. If detailed information is not available about ground albedo, it would appear reasonable for ground covered vegetation in temperate climates to adopt values of the order of 20–25%. Lower values should be used on central urban sites say 10–15%. Here building and road surfaces of lower basic albedo tend to dominate.

    Some of the limitations and difficulties of the method proposed are discussed in fuller detail in the original paper (1). A more refined approach requires that the non isotropic nature of the irradiation from the sky be properly considered. This is discussed fully later in this paper.

    8 EXAMPLE OF ESTIMATION OF MEAN IRRADIATION ON AN INCLINED SURFACE USING THE ISOTROPIC APPROXIMATION FOR DIFFUSE IRRADIATION

    o = 50%. Assuming a =.14 and b =.66 from ho = 22.5 MJm−2 day−1 from Table I we find:

    Adopting values of c =.94 and d = −1.03 from Table IV:

    bh = 5.8 MJm−2 day−1.

    For 45° slope, Ks =1.64

    Diffuse radiation on slope from sky = 4.8 × cos² 45/2 = 4.1 MJm−2 day−1

    Diffuse reflected radiation from ground (for Y =.25) = 10.6 x.25 × sin² 45/2 MJm−2 day−1 =.39 MJm−2 day−1. Mean monthly global radiation on a 45° slope, Gs=45° = (9.5 + 4.1 +.39) = 14.0 MJm−2 day−1.

    The ratio of the monthly mean irradiation on this inclined surface set at 45° to that on a horizontal surface for this date is thus estimated to be 14.0/10.7 or 1.31 compared with the clear sky direct component Ks value of 1.64. The mean amplification factor due to tilt for a cloudy climate is less than for a sunny climate.

    9 THE THEORETICAL EFFECT OF PERCENTAGE POSSIBLE SUNSHINE ON THE AMOUNT OF DIFFUSE RADIATION

    It is well known that the maximum amounts of diffuse radiation are received on days with partially clouded skies. Broken cumulus in particular can give very high values of diffuse radiation, and when the sun is shining between clouds, instantaneous values of the global irradiance can be recorded that are actually higher than the solar constant. One feature of the formula used for estimating diffuse radiation is that it mathematically predicts the increase of diffuse irradiation with intermediate amounts of sunshine. Since

    by rearranging equation 3 we get

    (7)

    This is a parabolic curve and one may find the maximum value by differentiation, i.e.

    For a maximum or minimum

    ho. This value corresponds to mid-day values of the horizontal diffuse irradiance of the order of 300 W/m² observed in lower latitudes with the sky 50% overcast.

    This analysis demonstrates the importance of proper consideration of diffuse irradiation in estimating the energy falling on inclined surfaces. However, to get greater accuracy, one has to consider in detail the effects of the non isotropic nature of the diffuse irradiation. Then the problem necessarily becomes more complicated. However, the very simple prediction method outlined has stood the test of time and is certainly useful in preliminary solar studies.

    10 THE NON ISOTROPIC NATURE OF THE IRRADIATION FROM THE SKY

    Equation 6 can only give the incoming monthly mean daily total irradiation on a slope correctly if the isotropic approximation is valid. This assumption must now be examined in more detail for it is well known that in practice the radiance of the sky is not isotropic. Unfortunately average diffuse irradiances on inclined surfaces have not been studied experimentally. Rather more attention has, however, been given to the study of the irradiance on inclined surfaces under cloudless conditions. This data will be reviewed first.

    11 DIFFUSE IRRADIATION ON CLEAR DAYS ON INCLINED SURFACES

    Among the earlier useful studies in this field may be mentioned the studies of Kondratyev and Manolova (5) in Russia and the work of Parmelee (6) in America. Parmelee’s work is confined to instantaneous values on vertical surfaces. Kondratyev’s work deals with instantaneous values on inclined surfaces as well. Parmelee’s work showed the great importance of atmospheric turbidity in determining the diffuse radiation on cloudless days. A decrease in the intensity of the direct beam is accompanied by an appreciable increase in the amount of diffuse radiation available on vertical surfaces. Kondratyev’s work showed that there are certain compensating effects which may make the isotropic approximation more reliable for average conditions than would at first sight appear. An important factor influencing the radiation balance of inclined surfaces is the change of albedo of the ground surface with the angle of incidence. Another important factor is that appreciably higher diffuse radiation is received on clear days from the quadrant of the sky facing the sun. The isotropic approximation is not satisfactory, overestimating the radiation on surfaces facing away from the sun, and underestimating the radiation on surfaces facing towards the sun and on surfaces at right angles to the sun’s azimuth as well on cloudless days and partially clouded days.

    12 ANALYSIS OF PARMELEE’S RESULTS FOR CLEAR DAYS

    Parmelee (6) presented detailed graphs showing the vertical surface irradiance for three atmospheric clarity conditions, defined in relation to Moon’s standard radiation curves, based on observations at Cleveland, USA. The clarity was expressed as a simple ratio Gbn/Gbn(MOON) where Gbn is the solar irradiance normal to the beam. As the corresponding values of the horizontal diffuse irradiance were also presented, it is possible, from Parmelee’s data, to calculate the ratio between vertical diffuse irradiance and the horizontal diffuse irradiance for different conditions of sky clarity. For an isotropic sky this ratio would, of course, be 0.5. Parmelee’s original observations included ground reflected irradiation and the ground albedo was stated to be 10%. One must subtract the predicted ground reflected radiation to get the diffuse vertical irradiance from the sky alone (Gds). Values of Gds/Gdh were calculated from Parmelee’s data and graphs of Gds/Gdh against cos i where i is the angle of incidence were prepared for different sky clarity conditions for different solar altitudes. The computed results for a Parmelee clarity of 0.8 are given in Figure 1. Figure 1 shows that for a sky clarity of 0.8 the vertical irradiance ratio observed by Parmelee was about three times the isotropic value for large positive values of cos i and less than half the isotropic value for large negative values of cos i. Clearly the isotropic approximation is not satisfactory for estimating vertical diffuse irradiation on cloudless days of different clarities and more accurate computational procedures are desirable.

    Fig. 1 Cloudless skies data derived from Parmelee.

    13 THE DEVELOPMENT OF AN IMPROVED CLEAR SKY SOLAR RADIATION MODEL

    The first prediction method described was based on the adoption of a standard radiation curve. In a more sophisticated approach proper account has to be taken of the effect of atmospheric turbidity on solar radiation. The first step in providing a clear sky short wave radiation model which could be used for surfaces of any slope and orientation anywhere on the earth’s surface was to adopt a suitable meteorologically validated radiation transmission model of the atmosphere which allowed for local variations in both atmospheric water content and sky clarity to predict the direct irradiance. The model proposed by Unsworth and Monteith (7) was adopted. This model had been rigorously checked out for different parts of the UK and similar checks for the Sudan had shown it to be reasonably applicable in lower latitudes also.

    The model is based on a turbidity coefficient, τa, which relates the measured direct normal solar irradiance Gbn to the clean air direct normal irradiance G*bn by the formula

    (8)

    where m is the optical air mass and G*bn is the irradiance normal to the beam for a clean atmosphere containing known amounts of absorbing and scattering gases and c is a correction factor to adjust for solar distance. The chief factors affecting G*bn, are the water vapour content of the atmosphere and the path length. The model adopted closely follows the theory described by Unsworth (8).

    The clear sky diffuse irradiance on a horizontal surface varies with both solar altitude and turbidity. Page (9) has shown that the diffuse irradiance measurements of Parmelee (6) compare favourably with measurements made in the UK and Belgium. Parmelee (6) presented observed results of the diffuse solar irradiance on horizontal and vertical surfaces. His results were derived from measurements made at Cleveland, Ohio on days which, although cloudless, differed considerably in atmospheric clarity. The horizontal surface results show that, for fixed solar altitude, a linear relationship exists between the diffuse irradiance Gdh

    Enjoying the preview?
    Page 1 of 1