Vous êtes sur la page 1sur 7

A Photovoltaic Module Thermal Model Using Observed Insolation and Meteorological Data to Support a Long Life, Highly Reliable

Module-Integrated Inverter Design by Predicting Expected Operating Temperature Robert S. Balog1


Senior Member, IEEE Texas A&M University 216P Zachary Engineering Center College Station, TX 77843, USA rbalog@ece.tamu.edu

Yingying Kuai
Student Member, IEEE University of Illinois, Urbana-Champaign 1406 W. Green St Urbana, IL61801, USA ykuai2@illinois.edu

Greg Uhrhan
SolarBridge Technologies 2111 S. Oak Street Champaign, IL 61821, USA uhrhan@yahoo.com

Abstract Accurate prediction of photovoltaic (PV) module temperature is needed to understand the expected electrical performance, lifetime, and reliability of photovoltaic cells. A photovoltaic AC module (PVAC) integrated the inverter directly with the PV module which exposes the power electronic circuitry to the thermal environment of the PV module. This has been reported to impose additional requirements on component selection and circuit design. However, a worst-case stack up analysis can lead to the conclusion that module-integrated inverters require industrial grade components or expensive thermal management. This paper presents a detailed thermal model for the PV module that uses real-world operating conditions, based on observed data from the National Renewable Energy Laboratory (NREL) to calculate PV module temperature. Results from the model confirm that the peak PV module temperature can reach over 80C, which was expected from other techniques, but that these peak temperatures occur on average for only 8 minutes per year in locations similar to Tucson, Az. Since the PV module temperature is found to be less than 70C for 99% of the operation hours, thermal management is not onerous and that the use of lower cost, commercial grade components will provide a mean time between failure (MTBF) to support an inverter warranty equivalent to that of the PV module itself. Index Terms Photovoltaic power systems, photovoltaic cell thermal factors, solar energy, solar power generation, thermal modeling.

I.

INTRODUCTION

Accurate prediction of photovoltaic (PV) module temperature is needed to understand the expected electrical performance, lifetime, and reliability of photovoltaic cells. Recent interest in integrating the power electronic inverter directly with the PV module [1, 2] means that the power electronics circuitry and components will now be exposed to the thermal environment on or near the PV module which can impose additional requirements on component selection and circuit design. Understanding the exact ambient thermal conditions expected both instantaneously as well as over the lifetime of the inverter is particularly important for circuit topologies that rely on aluminum electrolytic capacitors, which have well-known failure and degradation modes [3]
The author was a Senior Engineer at SolarBridge Technologies, 2111 S. Oak Street, Champaign, IL, USA, 61821 when the work in this paper was originally performed.
1

that are accelerated under these elevated thermal conditions found near a PV module. While other researchers have sought to improve the reliability of the power inverter by eliminating these electrolytic capacitors altogether [4, 5], the expected operating temperature due to the proximity of the electronics to the PV module can have an effect on other component selection and reliability. While it is tempting to apply a worst-case stack up analysis, this can lead to the conclusion that PV moduleintegrated inverters require industrial grade (105C) components [6]. This paper presents a thermal model for the PV module and simulation results that use real operating conditions, based on observed data from the National Renewable Energy Laboratory (NREL) [7] to calculate PV module temperature. Although the peak temperature was found to be 81C, the PV module temperature exceeded 80C for only 2 hours out of the 59,740 total operating hours in the 15 year dataset an average of 8 minutes per year. These results suggest that since the PV module temperature is less than 70C for 99% of the operation hours, good thermal management will support the use of lower cost commercial grade (70C) components. The results can also be used in calculating the thermal stress for mean time between failure predictions. The thermal model presented in this paper uses a controlvolume approach and takes into account incoming shortwave radiation (insolation), electrical conversion efficiency, longwave radiative exchange with the sky, earth, and roof, as well as free and forced convection from the top and bottom of the PV modules. Installation parameters include the roof pitch and the PV module tilt, which facilitates studying installation options other than parallel to the roof-line. The NREL database [7] provides actual meteorological data for 1,454 locations in the United States and is used with the model to generate a histogram of PV module operational temperatures. The model is designed for a roof-mounted PV module but is applicable to other mounting conditions, combined heat and power solar, and building-integrated solar applications. II. PREVIOUS WORK work in predicting PV module operating

Recent

978-1-4244-2893-9/09/$25.00 2009 IEEE

3343

Authorized licensed use limited to: AALTO UNIVERSITY. Downloaded on February 8, 2010 at 09:06 from IEEE Xplore. Restrictions apply.

temperature has applied a correlation approach where a known reference state, such as the normal operating cell temperature (NOCT) condition is extrapolated [8]. Since the NOCT is taken at a single operating point, this technique does not take into account heat loss that occurs under different sky, wind, and ambient conditions [9]. This paper extends the work in [10] which applies a first-principles based energy balance on the PV module control volume to solve for temperature. Accurate thermal models are important to understand and predict the performance of the silicon cells [11-13]. A recent trend has been to integrate the inverter electronics directly with the PV module [6]. Thermal stress limits the lifetime of critical components [3], overall mean time between failure (MTBF) and warranty programs [14]. III. MODELING APPROACH The thermal model treats the PV module as a total energy balance on the control volume [10]:

Table 1: PV material properties [10] PV module Element Density Specific Thickness of Layer, Heat dm cp [kg/m3] [J/kg-K] [m] 2330 1200 3000 677 1250 500 0.0003 0.0005 0.003 dmcp [J/m2-K] dmcp [WH/m2-K]

Silicon PV cells Polyester Tedlar Trilaminate Glass Face Total

473 750 4500 5723

0.13 0.21 1.25 1.59

assumed that losses from the edges are small compared to the total surface area. Further, the heat fluxes in (3), are assumed to be constant across the x-y plane of the PV module. Thus the PV module temperature can be represented as a single, lumped value for each instant in time and the governing differential equation (1) is a function only of time.
A. PV Module Heat Capacity In a standard transient thermal problem, stored energy, Q, is a function of the mass (m), specific heat (cp), and rate of temperature change (T) such that

qsw A Pout q loss A where


qsw , Pout , qloss , Ts , C , mod A,

dTs Cmod A = 0 dt

(1)

shortwave incident radiation W 2 m electrical power outout [W ] total heat flux out of the module W 2 m panel surface temperture [C ] heat capacity J 2 m K panel area m 2

Q = m c p T .

(5)

In the case of the PV panel, which is a multi-layer laminate, the equation can be generalized:

(2)

Q = A d m c p T ,

(6)

[ ]

The total heat flux out of the PV module is defined as:

where A is the PV module area, dm is the thickness of the laminate and and cp are composite values of density and specific heat, respectively. By considering the composition of each layer in the laminate, a composite heat capacity can be calculated:

q loss = q lw + q cond + qconvfree + qconvforced


where
q lw , q , cond qconvfree , qconvforced , longwave radiation loss W 2 m conductive heat loss W 2 m free convection heat loss W 2 m forced convection heat loss W 2 m

(3)
where

cmod = d mn n c pn , for n = 1...N layers ,


n

(7)

(4)

A discussion of each term in (1-4) follows in section IV. In general, the PV module temperature, Ts ( x, y, z , t ) , is a function of spatial location on the PV module in addition to time-varying environmental conditions. A 3-D solution to (1) could be found using numerical finite-element analysis methods, which would need to be solved for each time value of insolation and weather data. However, since the thickness of the PV module is small compared to the area, it can be

d , thickness of nth layer [m] mn density of nth layer kg 3 . (8) n , m c , specific heat of nth layer J kg K pn Table 1 provides an example calculation for a typical PV laminate [10] with material properties [15] representative of typical PV laminates. The PV module area, A, may be a free parameter such that all calculations are on a per unit area basis, allowing the model to be scaled to arbitrary sized installations, which is useful when multiple identical PV modules are mounted side-by-side to form large arrays.

NREL and NOAA Databases The input to the model is hourly insolation and meteorological data collected over the period of 1991- 2005 [7]. A minimum level of insolation is needed to generate

B.

3344

Authorized licensed use limited to: AALTO UNIVERSITY. Downloaded on February 8, 2010 at 09:06 from IEEE Xplore. Restrictions apply.

enough electrical power to power-up the inverter controls. A realistic tare value is 50W/m2 (approximately 5% of full illumination) which yields a theoretical 10W from a typical 200W, 1.5m2 , 14% efficient PV module. Thus, the analysis of the 15 year dataset reveals a total of 59,739 inverter operational hours, or a duty cycle of about 45%, for a location with insolation and weather similar to Tucson, Az. IV. DISCUSSION OF SPECIFIC TERMS PV Module Electrical Output Power: Pout The model supports arbitrary electrical conversion efficiency of incoming short wave radiation. Typical values for electrical conversion efficiency are 10% to 14% for commercially available multi-crystalline cells. The results in this paper assume a constant efficiency. In general, the efficiency is a function of temperature which adds a nonlinear term to the governing thermal equation (1). A. B. Long Wave Radiation: qlw Long wave radiation exchange is assumed to occur primarily between the PV module and the sky, earth and roof. The heat flux for the top side of the PV module is

The roof temperature can easily be 20C or greater than ambient temperature on a clear sunny day.

C.

Conductive Loss: qcond A thermal conduction path exists from the edge of the PV module, through the frame. Given the large surface area of the PV module and the small contact area of the frame, conductive heat flux can be assumed to be negligible.

Convective losses: qconv Convective heat transfer occurs due to free (natural buoyancy of hot air) convection qconvfree and forced convection qconvforced caused by wind. In standard heat transfer calculation, if one convective mode dominates, the other is usually ignored. In this model, both modes are currently taken into account at all times. The average temperature of the surface and ambient, Tavg, is used to calculate film properties:
(Ts + Tambient ) . 2 with a characteristic length of

D.

Tavg =

(13)

q lw _ top = g Fmst

g Fme

( (T

Ts4
4 s

4 Tsky

4 Tearth

)+ , )

L=

A . 2 * (H + W )

(14)

(9)

where is the Stefan-Boltzmann constant for radiative heat transfer and g is the emissivity of the module top surface, typically 0.9 to 1.0. View factors Fmst is from the module to the sky and Fme is from the module to the earth, both for the top surface. Tearth is the ground surface temperature, which is assumed to be the ambient air temperature for locations such as residential subdivisions with significant vegetation. The solar temperature is
4 Tsky = sky Tambient

The kinematic viscosity, , conductivity, k, and Prandtl number, Pr, are all determined over the range from 0C to 90C using curve-fitting software such as EES. Other coefficients needed to compute the Grasshof number are

=
and

1 Tavg

(15)

Tdiff = Ts Tambient .

(16)

0.25

(10)

The Grasshof number can then be computed using the standard definition GrL = 9.81 cos( ) Tdiff L3

where the emissivity of the sky, sky is determined by the dew point temperature Tdew [16]:

sky =

0.727 + 0.0060 Tdew , during daytime . (11) 0.741 + 0.0062 Tdew , during nighttime

(17)

where is the angle of the PV panel to the vertical. Then, the Rayleigh number is calculated by Ra L = GrL Pr . and a Function is tabulated as 0.492 16 9 ) ] . Pr The Nusselt number is then calculated as = [1 + ( N u L = 0.68 + 0.67 ( Ra L )0.25
9 16

where Tdew is in C and is obtained from the NREL database [7], while Tsky and Tambient are in Kelvin, [K]. Similarly, the heat flux for the bottom surface of the PV module is
4 q lw _ bot = b Fmsb Ts4 Tsky + 4 b Fmr Ts4 Troof

(18)

) )

(19)

(12)

where, b is the emissivity of the module bottom surface and is assumed to be equal to g. The roof temperature Troof is more complicated to compute since it is related to the ambient temperature, roofing material and construction, and solar heating of the roofing material near the PV module [15].

(20)

which yielding a free convection heat transfer coefficient of

h free =

NuL k . L

(21)

3345

Authorized licensed use limited to: AALTO UNIVERSITY. Downloaded on February 8, 2010 at 09:06 from IEEE Xplore. Restrictions apply.

In this model the heat transfer is assumed to be the same for both top and bottom surfaces, which is a fair assumption at near vertical angles but introduce error for PV modules mounted flat. A method of calculating heat transfer under forced convection, regardless of PV module orientation (inclination or yaw) can be found in [17], which is useful since the wind speed is a scalar value on a horizontal plate [7]. The specific heat and density product is found from a curve fit:
2 cprho = 1300.37 4.56864 T avg + .0116391 T avg (22)

From this an average heat transfer coefficient can be determined:

h forced ,t

wind cprho 0.931 L = 2

0.5

(23)

The data loading block allows the user to select system parameters such as location and PV panel type. The ODE solving block solves the above defined model and outputs PV module temperature as an hourly data array in C. Once the simulation has run for the entire data set the post-processing block performs statistical analysis on the data which includes generating histograms, cumulative distribution functions, and exceedence plots. Parameter inputs to the system comprise of insolation and meteorological data, PV module characteristic data, and installation parameters. Meteorological data included solar insolation, S, wind speed, WS, ambient temperature, Tambient, and dewpoint temperature, Tdew which are available from the NREL database [7] and represent hourly measured/estimated data for 15 years from 1991 to 2005. PV module data are obtained from manufacturer datasheets. Installation
Start Data Loading

Pr 3 Invoking Newtons Law of Cooling results in qconvforced = 2 h forced ,t Tdiff .

(24)

which assumes that both the top and bottom have the same heat transfer rates and are both exposed to the wind. In case of the panel being flush mounted with no air gap to the roof, the 2 factor would be eliminated. Free and forced convection are combined into an effective convective heat transfer coefficient:
3 3 Nu eff = Nu 3 forced Nu free

User Input: Location Selection

Load Insolation and Meteorological Data

User Input: PV module Selection

Load PV Module parameters

(25)

such that
heff = Nu eff k L

User Input: Installation parameters ODE Solver

Calculate view factors and mount angles

(26)

and
qconv = heff Tdiff .

Increment hour

(27)
Run thermal model

V. IMPLEMENTATION Rearranging (1) into a form suitable for simulation yields


dTs qsw A Pout q lw + q cond + qconv A = dt cmod A

NO

NO Solution Converged

(28)

which is a nonlinear, time-varying first-order ordinary differential equation (ODE) in terms of PV module temperature Ts since the terms qsw, qlw, and qconv are all functions of Ts and time t. MATLAB was used as the simulation environment and numerical solver. The Backward Euler method provided accurate results, was relatively fast to converge, and was stable whereas higher order methods, such as Runge-Kutta and Adams methods were more robust but took longer to simulate and did not provide significantly better results. The simulation contains three functional blocks as shown in Figure 1: data loading, ODE solving, and post-processing.

Last data point

Post Processing Post-processing statisitcal analysis and plotting

End

Figure 1: Simulation flowchart.

3346

Authorized licensed use limited to: AALTO UNIVERSITY. Downloaded on February 8, 2010 at 09:06 from IEEE Xplore. Restrictions apply.

Solar Insolation for 15 year(s) (NREL dataset: tucson ) 1000 [W/m ]


2

500 0

10

12 x 10
4

Ambient Temperature for 15 year(s) (NREL dataset: tucson ) 40 [ C] 20 0 2 4 6 8 10 12 x 10


[num ber of tenth] Sky cover for 15 year(s) (NREL dataset: tucson ) 10 5 0
4

10

12 x 10
4

Wind Speed for 15 year(s) (NREL dataset: tucson ) [m/s] 10 0

10
o

12 x 10
4

PV module temperature for 15 year(s) with Heat Capacity =1.59 (NREL dataset: tucsonRange: ) -12 ~ 79 C 60 40 20 0 2 4 6 Time (hour) 8 10 12 x 10
4

[ C]

Figure 2: Time trace for temperature and weather data for Tucson, Az.

parameters are set by the user and can accommodate correction for geographic location, time-of-day peaking, as well as roof pitch and orientation. VI. RESULTS Simulations have been performed for a wide variety of sites throughout the United States. The following set of figures illustrates the type of input and output data for the model. The results are based on a typical commercially available 195W PV module installed on a roof a 14.0 pitch and a PV module tilt of 14.3. The model allows for these installation parameters to be arbitrarily varied. Figure 2 shows the time traces for the simulation input data and the resulting PV module temperature. Figure 3 shows the histogram of the PV module temperature during the time the inverter is delivering power. In practice, a minimum level of insolation is needed to generate sufficient electrical power for the inverter controls, which results an inverter duty cycle of about 45% for a location with insolation and weather similar to Tucson, Az The overall shape of the curve reveals that the PV module temperature is less that 60C for more than 95% of the time

and the cumulative distribution function (CDF), in Figure 4, shows that the PV module temperature is less than 70C for 99% of operation hours. Although the peak PV module temperature was found to be 81C, the PV module temperature exceeded 80C for only 2 hours out of the 59,740 total operating hours as shown in Figure 5. The CDF and exceedence plots reveal that the module-integrated inverter need not be designed for continuous duty at or above 80C. Using the temperature histogram the design engineer can compute a composite MTBF, which may be a weighted average of the MTBF at a number of different temperatures, which will reflect actual expected operation. The PV module temperature is affected by different installation methods such as if mounted close to the roof which blocks airflow on the bottom side or on a rack system with unimpeded air flow. Further, different roofing materials also affect the module temperature. All of these installationspecific variations may be considered simultaneously, individually, or by using a statistical technique such as MonteCarlo analysis to determine the design margin sufficient for an intended application. When multiple geographic regions are compared, the data reveals that the hottest ambient air temperature does not

3347

Authorized licensed use limited to: AALTO UNIVERSITY. Downloaded on February 8, 2010 at 09:06 from IEEE Xplore. Restrictions apply.

Histogram of PV panel temperature for 15 years during inverter operation - 59740 operating hours 3 Percentage over operating hours [%] 2.5 2 1.5 1 0.5 0 -20

-10

10

20

30

40

50

60

70

80

90

Temperature [oC]
Figure 3: Example of a PV module temperature histogram for inverter operational hours.
Cum ulative Dis tribution Function 1 PV panel temperature temperature

High Tem perature 100

Percentage of total hours

0.8

80 Number of hours
0 20 40
o

0.6

60

0.4

40

0.2

20

0 -20

60

80

0 74

76

78

80

82

Temperature [ C]

Exceedence Temperature [oC]

Figure 4: Example of a cumulative distribution function for PV module temperature.

Figure 5: Example of an exceedence temperate plot.

always result in the hottest PV module temperature. This is significant since many empirical system-design rules predict worst-case PV module temperature by adding a constant offset to the worst-case ambient air temperature. However, sky conditions and wind speed are not considered in these approaches yet the model shows that they have a significant impact on module temperature. Results from the thermal model and simulation also can provide daily temperature cycle data, as shown in Figure 6. Diurnal thermal cycling can have a significant impact on inverter reliability due to mis-matched coefficient of thermal expansion as well as stress and fatigue on mounting hardware, connectors, and solder connections. Comparing diurnal PV module temperature cycles in different geographic regions revealed that the region with the worst-case peak module temperature did not necessarily have the greater daily thermal cycle. VII. CONCLUSION

Accurate prediction of PV module temperature is important to understand performance, reliability and lifetime of a PV module-integrated inverter. Whereas other techniques correlate and extend a single measured operating point, this paper presents a first-principles thermal model and simulation methodology that uses an energy-balance approach along with observed insolation and meteorological data to compute the temperature of the PV module using historical, measured data. The result is an hour-by-hour PV module temperature which is post-processed to produce a histogram of temperature distribution, cumulative-distribution function, and exceedence plot. Additional post-processing on the PV module temperature data can produce distributions of expected diurnal temperature cycles. The complete set of governing thermal equations needed to directly implement the model has been included. The data provided from this thermal model can be used by design engineers to optimize the component selection, cost and thermal management of the power electronic inverter in

3348

Authorized licensed use limited to: AALTO UNIVERSITY. Downloaded on February 8, 2010 at 09:06 from IEEE Xplore. Restrictions apply.

PV Panel Daily Temperature Cycles for 15 years (NREL dataset: tucson) 25

20 Percen tage of cycles fo r 15 yea rs

15

10

-5

10

15

20

25

30

35
o

40 [ C]

45

50

55

60

65

Temperature Difference | T|

Figure 6: Example of PV module daily temperature cycles.

order to meet cost and reliability objective and achieve the long lifetime and high reliability needed from a PV moduleintegrated inverter. REFERENCES
[1] Q. Li and P. Wolfs, "A Review of the Single Phase Photovoltaic Module Integrated Converter Topologies With Three Different DC Link Configurations," IEEE Transactions on Power Electronics, vol. 23, no. 3, pp. 1320-1333, May 2008. S. B. Kjaer, J. K. Pedersen, and F. Blaabjerg, "A review of single-phase grid-connected inverters for photovoltaic modules," IEEE Transactions on Industry Applications, vol. 41, no. 5, pp. 1292-1306, Sept.-Oct. 2005. J. L. Stevens, J. S. Shaffer, and J. T. Vandenham, "The service life of large aluminum electrolytic capacitors: effects of construction and application," IEEE Transactions on Industry Applications, vol. 38, no. 5, pp. 1441-1446, Sept.-Oct. 2002. P. T. Krein and R. S. Balog, "Cost-Effective Hundred-Year Life for Single-Phase Inverters and Rectifiers in Solar and LED Lighting Applications Based on Minimum Capacitance Requirements and a Ripple Power Port," in Applied Power Electronics Conference and Exposition, 2009. APEC 2009. Twenty-Fourth Annual IEEE, pp. 620625. A. C. Kyritsis, N. P. Papanikolaou, and E. C. Tatakis, "A novel Parallel Active Filter for Current Pulsation Smoothing on single stage gridconnected AC-PV modules," in Proceedings, IEEE European Conference on Power Electronics and Applications, 2007, pp. 1-10. B. Sahan, N. Henze, A. Engler, et al., "System design of compact lowpower inverters for the application in photovoltaic AC-modules," in 5th International Conference on Integrated Power Electronic Systems, CIPS 2008, VDE Verlag GmbH, Berlin, Germany, March 2008, pp. 187-92. NREL, "National Solar Radiation Data Base (NSRDB) 1991- 2005 Update," National Renewable Energy Laboratory (NREL). Available: http://rredc.nrel.gov/solar/old_data/nsrdb/1991-2005/. E. Skoplaki and J. A. Palyvos, "Operating temperature of photovoltaic modules: a survey of pertinent correlations," Renewable Energy, vol. 34, no. 1, pp. 23-29, 2009. E. Skoplaki, A. G. Boudouvis, and J. A. Palyvos, "A simple correlation for the operating temperature of photovoltaic modules of arbitrary

[10] [11] [12]

[2]

[13]

[3]

[14] [15] [16] [17]

[4]

mounting," Solar Energy Materials and Solar Cells, vol. 92, no. 11, pp. 1393-1402, Nov 2008. A. D. Jones and C. P. Underwood, "A thermal model for photovoltaic systems," Solar Energy, vol. 70, no. 4, pp. 349-359, Mar 8, 2001. K. Emery, J. Burdick, Y. Caiyem, et al., "Temperature dependence of photovoltaic cells, modules, and systems," in Record, IEEE Photovoltaic Specialists Conference, 1996, pp. 1275-1278. A. H. Fanney, M. W. Davis, B. P. Dougherty, et al., "Comparison of photovoltaic module performance measurements," Transactions of the ASME. Journal of Solar Energy Engineering, vol. 128, no. 2, pp. 1529, May 2006. G. Notton, M. Mattei, C. Cristofari, et al., "Calculation of the polycrystalline PV module temperature using a simple method of energy balance," Renewable Energy, vol. 31, no. 4, pp. 553-67, Apr. 2006. S. Vittal and R. Phillips, "Modeling and Optimization of Extended Warranties Using Probabilistic Design," in Proceedings, IEEE Reliability and Maintainability Symposium, 2007, pp. 41-47. "ASHRAE Fundamentals Handbook." Atlanta, Ga.: American Society of Heating Refrigerating and Air-Conditioning Engineers, 2001. A. F. Mills, Heat and mass transfer. Burr Ridge, Ill.: Irwin Inc., 1995. E. M. Sparrow and K. K. Tien, "Forced convection heat transfer at an inclined and yawed square plate-application to solar collectors," Transactions of the ASME, Journal of Heat Transfer, vol. 99, no. 4, pp. 507-12, Nov 1977.

[5]

[6]

[7] [8] [9]

3349

Authorized licensed use limited to: AALTO UNIVERSITY. Downloaded on February 8, 2010 at 09:06 from IEEE Xplore. Restrictions apply.

Vous aimerez peut-être aussi