You are on page 1of 28

Accepted Manuscript

The integration of Semi-transparent photovoltaics on greenhouse roof for energy


and plant production

Reda Hassanien Emam Hassanien, Ming Li, Fang Yin

PII: S0960-1481(18)30044-2

DOI: 10.1016/j.renene.2018.01.044

Reference: RENE 9649

To appear in: Renewable Energy

Received Date: 04 July 2017

Revised Date: 07 January 2018

Accepted Date: 14 January 2018

Please cite this article as: Reda Hassanien Emam Hassanien, Ming Li, Fang Yin, The integration of
Semi-transparent photovoltaics on greenhouse roof for energy and plant production, Renewable
Energy (2018), doi: 10.1016/j.renene.2018.01.044

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form.
Please note that during the production process errors may be discovered which could affect the
content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

1 The integration of Semi-transparent photovoltaics on greenhouse roof for


2 energy and plant production
3
4 Reda Hassanien Emam HASSANIEN* 1, 2, Ming LI1*, Fang YIN1
5
6 1 Solar Energy Research Institute, Yunnan Normal University, Kunming 650500, China.

7 2 Agricultural Engineering Department, Faculty of Agriculture, Cairo University, Giza 12613, Egypt.

8
9 Abstract
10 The aim of this study was to investigate the effect of semi-transparent building integrated
11 photovoltaics (BIPV) mounted on top of a greenhouse, on the growth of tomatoes and
12 microclimate conditions as well as to estimate the generated energy and the payback period of
13 this system. Three modules were settled at 20 % of the greenhouse roof area at a tilt angle of 30º
14 facing south at a distance of 0.08 m between the plastic cover and the BIPV. Each module has a
15 peak power of 170 Wp and efficiency of 8.25 %. Results revealed that the annual generated
16 electric energy of the BIPV was 637 kWh. Furthermore, there were no significant differences
17 (P< 0.05) in the growth of tomatoes between shaded greenhouse by the BIPV and the un-shaded
18 greenhouse. The reduction of solar radiation under the BIPV was 35 % - 40 % more than the
19 Polyethylene covers on clear days. The BIPV shading decreases the air temperature by (1 ºC - 3
20 ºC) on clear days and has no effect on relative humidity. The payback period was found to be 9
21 years. Moreover, this system can provide most of the annual energy demands for the greenhouse
22 environmental control systems.
23
24 Key words: Semi-transparent BIPV photovoltaics, greenhouse, energy, shading, tomato
25
26 1*Corresponding authors: Reda Hassanien & Ming LI
27 Tel./fax: +86 871 65517266.
28 E-mail address: reda.emam@agr.cu.edu.eg
29 E-mail address: lmllldy@126.com
30
31
32
33
34

1
ACCEPTED MANUSCRIPT

1
2 1. Introduction
3 Projections for population growth and food demand through 2050 result in a clear increase of
4 food and energy demands [1]. In addition, the continuously increasing scarcity of conventional
5 fuel, particularly fossil fuel resources, the need for greenhouse gas emissions reduction and the
6 global climate changes call for urgent actions to implement more sustainable technologies to
7 move towards climate-smart agriculture for food security and energy production [2].
8 Consequently, photovoltaics panels (PV) are an attractive renewable energy technology because
9 they avoid significant carbon emissions during their usage, have a long useful lifetime estimated
10 at 20-30 years, and they take advantage of a stable and plentiful energy resource [3]. Djevic and
11 Dimitrijevic [4] studied the influence of greenhouse construction on energy consumption in
12 winter lettuce production for four greenhouses in the Serbia region. They reported that the multi-
13 span greenhouse had the lowest energy consumption (i.e., 2.71 kWh/m2). The energy use pattern
14 for tomato production in greenhouses indicated that diesel, electricity and chemical fertilizers are
15 the major energy consuming inputs in Iran [5] and the energy consumption (direct and indirect)
16 per unit floor area of greenhouse in Indonesia for tomatoes was 47.62, GJ/ha [6]. Canakci and
17 Akinci [7] also reported that tomatoes cultivation is the most profitable crop among the
18 greenhouse vegetables in Turkey. Recently, a number of studies have been conducted on the
19 integration of PV panels on the agricultural greenhouses [8-11]. Integration of PV with cropland
20 can partially decrease the water consumption in irrigation [12, 13], alleviate the increasing
21 competition for land between food and energy production [14], reduce, or replace part of the
22 electrical energy consumption [15, 16] and decrease the greenhouse gas emission by 243,252
23 tons per year on Caribbean islands [17]. Additionally, it has been reported that PV module prices
24 have been reduced in the past decade by 80 % [18]. Furthermore, the costs of off-grid hybrid PV
25 systems were 19 % cheaper compared with electricity generation by diesel generators in most
26 rural parts of Indonesia, whereas stand-alone PV systems were 3 % cheaper than stand-alone
27 diesel generators on average [19]. Kadowaki et al.[10] found that the straight-line (PVs) array
28 shading significantly decreased the fresh weight and dry-matter weight of welsh onions
29 compared to the checkerboard (PVc) and control as well as decreased the annual crop yield by
30 25 % in Japan. The effect of flexible solar panels mounted on top of a greenhouse for electricity
31 production on yield and quality of tomatoes has been investigated in southeastern Spain [15].

2
ACCEPTED MANUSCRIPT

1 Results revealed that there was no effect found in yield and price of tomatoes despite their
2 negative effect on the fruits size and color. Cossu et al. [20] introduced a novel algorithm to
3 estimate the cumulated global radiation (GGR) inside photovoltaic (PV) greenhouses. They
4 found that the yearly GGR increased with the canopy height on the zones under the plastic cover
5 from 59 % at 0.0 m to 73 % at 2.0 m, and decreased under 50 % PV cover ratio on the roof from
6 57 % at 0.0 m to 40 % at 2.0 m. Fatnassi et al. [21] predicted the distributed microclimate inside
7 greenhouse equipped with photovoltaic panels by the Computational Fluid Dynamic (CFD)
8 model. They reported that the mean solar radiation transmission in asymmetric greenhouse,
9 which covered an area of 1 ha, composed of six Asymmetrical modules, and equipped with a six
10 continuous openings on the roof, was 41.6 % whereas that of the Venlo greenhouse, which had a
11 10 Venlo type module, covered an area of 1.4 ha, equipped with three continuous openings in the
12 roof, was 46 %. On the other hand, the integration of the BIPV with buildings can serve as a
13 shading device for a window, a semi-transparent glass facade, a building exterior cladding panel,
14 a skylight, and parapet unit or roofing system [22]. Li et al. [23] reported that the semi-
15 transparent PV facade in the sub-tropical climate of Hong Kong was able to reduce the annual
16 building energy use and peak cooling load by 1203 MWh and 450 kW, respectively for a room
17 located at the 20th floor of a residential building facing close to west (260º). Moreover, BIPV can
18 generate electricity at a building’s peak usage times and reduce the building’s peak grid
19 electricity demand [24]. Meanwhile, the heat from solar radiation in the air channel between the
20 BIPV panels or glazing and the building façade could be helpful to decrease the heating load in
21 winter at temperate and cold climates [25, 26].

22 The environmental control in summer is one of the major challenges faced by greenhouse
23 growers in the tropical and sub-tropical climates. Thus, shading the agricultural greenhouses in
24 hot and sunny regions along with cooling systems reduces water consumption by 25 %, reduces
25 greenhouse air temperatures below the outside temperatures by 5-10° C and increases the relative
26 humidity by 15 %-20 %. Moreover, shading reduces the solar radiation by 30 %-50 %, and
27 decreases the energy consumption for cooling by 20 % and 15 % for heating [27]. Li et al.[28]
28 reported that the annual return of the integrated photovoltaic and agricultural greenhouses in
29 China was varied from 9 % to 20 % with a payback period of 4-8 years. The previous studies
30 [29-33] which conducted in the integration of PV array on greenhouses commonly used
31 conventional opaque PV or planar flexible PV modules directly above the roof or under the roof.

3
ACCEPTED MANUSCRIPT

1 However, the internal integration (under the plastic cover) could decrease the efficiency and
2 lifespan of BIPV panel due to the high humidity rates, air temperatures and the agrochemicals
3 inside the greenhouses. Meanwhile, the high reduction of solar radiation may affect the plant
4 growth and lead to pathogen development [21]. In addition, no experimental data reported to date
5 has shown how tomato plants grow under semi-transparent BIPV shading conditions. Therefore,
6 the aim of this study is to find out the effect of semi-transparent BIPV panels on the growth of
7 tomatoes (Solanum lycopersicum L., cv. cherry). Furthermore, to estimate the generated
8 electrical energy at different tilted angles and areas of the roof and the payback period for
9 achieving a sustainable greenhouse crop production.

10 2. Materials and Methods


11 2.1. The greenhouse structure and PV configuration
12 This experiment has been conducted in Kunming, China (longitude 102.68°E and latitude
13 25.07°N), in two greenhouses of a single stand-alone structure with an equal gable roof in east-
14 west orientation to maximize the utilization of solar energy on the south roof in winter. The
15 dimensions of each greenhouse were 7.5 m (length), 3.5 m (width), 3 m (maximum height) and 2
16 m (gutter height), with a roof slope of 30°. The material covering the greenhouses was 0.12 mm
17 plastic Polyethylene (PE) film with light transmittance of 80 %. The ventilation vents were along
18 the side walls (7 m×1m) with net curtains of white saran. Ventilation was provided by manually
19 opened side windows. The first greenhouse was considered as shading treatment with semi-
20 transparent BIPV and the second greenhouse with Polyethylene cover only was as a control.
21 Three semi-transparent BIPV of mono-crystalline silicon double glazed were fixed on the south
22 west roof of one greenhouse at tilted angle of 30 and settled on top of the plastic cover with a
23 vertical distance of 0.08 m as an isolation air channel between the plastic cover and the BIPV to
24 reduce the heat stress and to avoid the high internal air temperatures and relative humidity from
25 the greenhouse to the modules as well as from the BIPV to the greenhouse. Each module (1985
26 mm×1038 mm) had a peak power of 170 Wp and efficiency of 8.25 %. The transmittance ratio of
27 each module was 47 % (64 cells, 12 cm ×12 cm) and the total area was 6.1 m2, which occupied
28 20 % of the greenhouse roof area as shown in Fig.1. The peak rated power of the three modules
29 was 510 Wp. The electrical and mechanical characteristics of the BIPV and Micro-inverters are
30 shown in Table1.
31

4
ACCEPTED MANUSCRIPT

1 Table 1 Specifications of the solar panels and the DC/AC inverter system.
Solar panels DC/AC Micro inverter system
Name and type CD-BIPV-64/5M170 Name YC250
AC Power(max) 250 W
Peak power (Wp) 170 W AC Vol 230-270 VAC
tage
Peak voltage (V) 33.2 V Frequency 50-55 HZ

Peak current (A) 5.12 A Max. Efficiency 95.5 %

Open circuit voltage (V) 39.6 V

Short circuit current (A) 5.65 A

Layer structure Toughened glass(6mm)+PVB(2.28mm)+(125

×125mm) sc-Si cells+ toughened glass(6mm)


Weight per piece (kg) 63 kg
2

(a) th (b)
Sou

(c) (d)

3 Fig.1 Photo of the greenhouses (a): Shaded greenhouse: [1] pyranometer, [2] Recording station, [3] Side ventilation
4 system, [4] The distance between the BIPV panels and plastic cover, [5] Micro-inverter’s cable. (b): Un-shaded
5 greenhouse, (c): BIPV panel, (d): greenhouse diagram.

6 The environmental control parameters such as air temperature, solar radiation, and relative
7 humidity were measured both outside and inside the greenhouse. The greenhouse roof of 30 m2
8 area, had a supporting structure consisted of 7 north-south oriented steel pipes of 0.03 m
9 thicknesses, which shaped gable roof. These were mutually separated by 1.25 m. The total

5
ACCEPTED MANUSCRIPT

1 coverage area of the pipes on the roof was 3 m2, which occupied 10 % of the roof area. Two
2 white light-emitting Diode (LED 30 W, China) lamps were attached to horizontal steel bars at a
3 height of 2 m, providing light intensity of 2700 lm. A small circulation fan with a discharge of
4 3000 m3/h was fixed in the center of the greenhouse (0. 370 kW and 220V).
5 The greenhouse transmissivity was calculated as the ratio between the internal global radiation
6 measured under the plastic cover and the external global radiation [15]. The solar panels were
7 connected in a series and fed into the grid by a three DC/AC Micro-inverters APS YC250A,
8 (Zhejiang Yu energy technology, Limited, China).The energy consumed by the greenhouse (e.g.,
9 for ventilation and lighting) and the AC real power output of the Micro-inverters were also
10 measured using smart electric meters DrF-AVC-485 (Beijing Hanwei Borch technology Limited,
11 China) and smartPV software at 4 s intervals. In addition, PolySun program was used to simulate
12 and predict the annual generated energy at different angles and covering area of the greenhouse
13 roof. The simulation of the BIPV was done before the installation. However, the tilt angle was
14 chosen to be 30º to fit the length of the BIPV panel with the greenhouse structure and the
15 experiment was conducted to cover only part of the south roof to save the initial cost of the
16 experiment. Subsequently, the cost of the BIPV panels and the payback period were calculated in
17 terms of energy self-provision and the income from the sale of the remaining energy to the grid.

18 2.2. Plants measurements


19 Tomato seedlings (Solanum lycopersicum L., cv. cherry) were transplanted in pots at plant
20 spacing in each row of 0.5 m and distance between rows of 1 m. There were 42 pots for each
21 greenhouse, each pot housing one plant. All plants had the same soil, irrigation and fertilization
22 systems. The experimental design was a randomized block design with two treatments (shading
23 and un-shading greenhouses) and four replicates in each greenhouse. The plant measurements
24 were plant height, number of leaves, leaves area, chlorophyll contents on 37, 47, 67 and 87 days
25 after shading treatments. The relative chlorophyll concentration of the leaves was measured by a
26 portable chlorophyll meter (or SPAD meter) and leaf area by a portable leaf area meter, both
27 instruments from Shanghai Rong Yan instruments. Co., Ltd, China. Four different experiments
28 with four different plants (such as lettuce, strawberry, cucumber, and tomato) have been
29 conducted in these integrated BIPV greenhouses. However, only results for tomato plants were
30 presented in this manuscript. The experimental data were analyzed using the statistical software
31 SPSS 18.0 (SPSS Inc., Chicago, IL, USA). Data and results from each sampling were analyzed

6
ACCEPTED MANUSCRIPT

1 separately by one-way analysis of variance (ANOVA). The data were given as the mean ±
2 Standard errors and the level of statistical significance was set at P < 0.05.
3 2.3. Microclimate monitoring
4 Mean daily relative humidity and air temperature were measured. Relative humidity was
5 recorded each hour by an Accurate TH12R-EX recorder from Xin yada instrument Co., Ltd,
6 China. The light intensity was measured by a digital portable lux meter from Taiwan, Tai Pei
7 TES electronic industry. Ltd., China. The solar radiation at crop level was measured each minute
8 in each treatment with pyranometers sensors connected to a data logger (model TRM-2,
9 Beijing Tianyu technology, Ltd., China). Two sensors for solar radiation were installed at two
10 levels of plants and fixed in the southwest under the BIPV and in the middle of the un-shaded
11 greenhouse at the same height. Two sensors for air temperatures were distributed in each
12 greenhouse at 1.0 m and 1.5 m above the ground level and at 1.5 m apart as shown in Fig. 2. The
13 specifications of all sensors are shown in Table 2 and the monthly averaged weather variables in
14 Kunming are shown in Table 3.

Shaded Gh Unshaded Gh

1300
3500

W E

1500
7500

15
16 Fig.2. Sensors position and distributions. This bullet represent the solar radiation, light intensity and relative
17 humidity sensors. Meanwhile, this bullet represent air temperature sensors in shaded and un-shaded greenhouses
18 (units in mm).

19

20

21

22

23

7
ACCEPTED MANUSCRIPT

1 Table 2.The specification of the instruments

instrument name Measurement range Accuracy

Thermocouple 0 - 150 ℃ ± 0.1℃

Pyranometers 0 - 2000 W/m2 ±2%

Wind speed sensor 0 - 70 m/s 0.5 m/s

Digital portable lux meter (0-200000 lx). ±1lx

Accurate TH12R-EX 0 - 100 % ± 0.2 ℃ and ± 2 %

3 Table.3.Monthly-averaged of meteorological data for Kunming city[34]

Months Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

Average temperature(℃) 5.8 7.9 12.2 16.4 18.2 19.3 19.2 18.9 16.6 13.0 9.41 6.20

Average relative humidity (%) 75.8 66.9 55.8 50.6 60.5 71.7 74.9 75.5 77.9 82.3 81.2 80.4

Daily solar irradiation - horizontal


4.38 5.13 5.71 6.23 5.61 5.04 4.57 4.59 4.03 3.77 3.92 3.97
(kWh/m2/d)

4
5 3. Results and discussions
6 3.1. Effect of BIPVs on the greenhouse air temperature and relative humidity
7 Climate distribution in greenhouses can greatly influence the growth and development of crops,
8 particularly the air temperature distributions. It was observed that air temperature under the
9 shaded greenhouse by BIPV was lower than that of the un-shaded greenhouse during the period
10 from 12:00 noon to 2: 00 p.m. in the range of (1 ºC – 3 ºC). Consequently, the average
11 temperature differences between the shaded and un-shaded greenhouses were high only on sunny
12 days and the variation started from 9 a.m. to 3 p.m. The average air temperature inside the
13 shaded greenhouse with natural ventilation was higher than that of the ambient air temperature
14 by (5 ºC -8 ºC) and the average air temperature inside the un-shaded greenhouse was higher than
15 the ambient temperature by (8 ºC-10 ºC) as shown in Fig.3a. Furthermore, the average air
16 temperature in the air channel between the plastic cover and the BIPV on top of the greenhouse
17 roof was lower than the internal air temperature of greenhouse by (3 ºC - 6 ºC) on sunny days
18 meanwhile, higher than the ambient air temperature by (3 ºC -5 ºC) as shown in Fig.3b.

8
ACCEPTED MANUSCRIPT

45
Shaded Gh Un-shaded Gh Ambient temperature
40

35
Air temperature (ºC)

30

25

20

15

10

0
0:01
10:34
13:34
6:01
16:34
19:34
22:34
10:34
1:34
13:34
4:34
16:34
7:34
19:34
22:34
3:09
15:09
9:09
21:09 23:58 23:58 23:58 23:58 23:58 23:58

Time ( h)

(a) 10th- 16th May,2017


Shaded Gh Between the BIPV and the plastic cover Ambient temperature

41.0
38.0
35.0
32.0
Air temperature (ºC)

29.0
26.0
23.0
20.0
17.0
14.0
11.0
8.0
5.0
7:00 8:00 9:00 10:00 11:00 12:00 13:00 14:00 15:00 16:00 17:00 18:00 19:00

Time (h)

(b) 26th April, 2016

1 Fig.3 Average air temperature inside and outside the greenhouses under natural ventilation.
2 Figure 4 shows that the temperature of the BIPV was higher than the internal air temperatures of
3 the greenhouse by an average of 8 ºC, while the difference between the module temperature and
4 the ambient temperature was 14 ºC. The maximum and minimum temperatures of the BIPV were
5 48.3 ºC and 23 ºC, respectively while the maximum and minimum temperatures of the

9
ACCEPTED MANUSCRIPT

1 greenhouse under the natural ventilation were 38 ºC and 23.3 ºC, respectively. These results
2 illustrated that the integration of the BIPV outside the greenhouse could alleviate the over-
3 heating inside the greenhouse in summer.

BIPV Inside the shaded Gh Ambient temperatures


55.0

50.0

45.0

40.0
Air temperature (ºC)

35.0

30.0

25.0

20.0

15.0

10.0

5.0

0.0
8:00 9:00 10:00 11:00 12:00 13:00 14:00 15:00 16:00 17:00 18:00 19:00 20:00

Time (h)
4
5 Fig. 4 Average air temperatures inside and outside the shaded greenhouse compared to the BIPV temperature on 10th
6 July, 2016.
7 Data here in Figures 3 and 4 were presented to show the shading effect on a certain sunny days
8 of hot months from April to July, due to the fact that there were obvious differences in air
9 temperatures during those months. Results revealed also that air temperature decreased in the un-
10 shaded greenhouse compared to the shaded greenhouse from 5 p.m. to 7 p.m. as shown earlier in
11 Fig3a, which means that the un-shaded greenhouse was losing heat faster comparing to shaded
12 greenhouse due to the lower solar radiation in late afternoon. Furthermore, integrating the BIPV
13 on the greenhouse could decrease the heat losses at night in winter due to the BIPV and air
14 channel between the plastic cover and the BIPV could serve as a thermal cover to maintain the
15 internal air temperature. In contrast, it decreases the internal air temperatures during the daytime
16 in summer, these results are in agreement with Ahmed et el.[27].

10
ACCEPTED MANUSCRIPT

1 Air temperatures inside greenhouses were always observed higher than that of the ambient
2 temperatures. In contrast, the relative humidity was lower inside the greenhouse on yearly basis
3 particularly during the day time on sunny days. It was clear that using the natural ventilation all
4 day long could decrease the internal air temperature and relative humidity however, on rainy
5 days the relative humidity increased to 99.9 %. Thus, the mechanical ventilation operated only
6 on hot days when the internal temperature exceeded 25 ºC and the relative humidity 85%. It was
7 observed that relative humidity increased gradually at night from 6:00 p.m. to 6:00 a.m. inside
8 the greenhouse compared to the outside. The average relative humidity in the shaded greenhouse
9 was insignificantly higher than that of the un-shaded greenhouse by 2 % during the day in
10 summer. The maximum internal relative humidity was 99.9 at night and on rainy days and the
11 minimum was approximately 23 % in the un-shaded and 25 % in the shaded greenhouse during
12 the daytime particularly afternoon on sunny days at high dry bulb temperatures as shown in
13 Figs.5a and 5b.

110.0
Shaded Gh Un-shaded Gh Outside
100.0

90.0
Relative humidity(%)

80.0

70.0

60.0

50.0

40.0

30.0

20.0

10.0
2016-07-21 2016-07-21 2016-07-22 2016-07-23 2016-07-24 2016-07-25 2016-07-26 2016-07-26
00:33:17 20:33:17 16:33:17 12:33:17 08:33:17 04:33:17 00:33:17 20:33:17

Date and time

(a) Relative humidity

11
ACCEPTED MANUSCRIPT

45
Shaded Gh Outside
40
Dry bulb temperature (ºC)

35

30

25

20

15

10

5
2016-07-21 2016-07-21 2016-07-22 2016-07-23 2016-07-24 2016-07-25 2016-07-26 2016-07-26
00:03:01 20:03:01 16:03:01 12:03:01 08:03:01 04:03:01 00:03:01 20:03:01

Date and time

(b) Dry bulb temperature


1 Fig.5 Relative humidity and dry bulb temperature inside and outside the greenhouses under the natural ventilation.
2 3.2. Effect of semi-transparent PV on solar radiation
3 Results showed that the solar radiation under the shaded greenhouse by the plastic cover of
4 Polyethylene (PE) along with BIPV panels was lower than that of PE cover only. Accordingly,
5 the plants under the PE cover always received a higher amount of solar radiation compared to
6 those under the BIPV with PE cover.
7 Figure 6a shows the results of a primary experiment which was conducted in the shaded
8 greenhouse only to compare between the shaded (west) and un-shaded (east) parts. It was
9 observed that the solar radiation in the morning was lower under the shaded part. Subsequently,
10 the variation in the solar radiation changed at 3:00 p.m. due to the shaded area moves in the same
11 greenhouse from the west to the east direction in terms of time and the solar radiation sensors
12 were in a fixed position. Nevertheless, there was an extra shading from 11:00 a.m. to 1:00 p.m.
13 which caused by the greenhouse structure, due the shading of greenhouse structure was very
14 difficult to be avoided. Subsequently, it was found that solar radiation in the shaded greenhouse
15 increased from 1:30 p.m. to 2:45 p.m. due to the movement of most of the shaded area away
16 from the solar radiation sensor during this period (Fig.6b). Meanwhile, the solar radiation
17 decreased in the un-shaded greenhouse at 1:00 p.m. as a result of the greenhouse structure.

12
ACCEPTED MANUSCRIPT

1200
Outside Shaded part with PV Un-shaded part

1000
Solar radiation (W /m2)

800

600

400

200

0
6:58 7:58 8:58 9:58 10:58 11:58 12:58 13:58 14:58 15:58 16:58 17:58 18:58
Time (h)

(a) 26th May 2015, inside the shaded greenhouse (plastic and plastic with the BIPV
panels) and outside the greenhouse at a 0.60 m from ground
1
1400
Un-shaded Outside Shaded
1200
Solar radiation (W /m2)

1000

800

600

400

200

0
7:07 8:09 9:11 10:07 11:09 12:11 13:07 14:09 15:11 16:07 17:09 18:11
Time (h)

(b) 22nd May, 2017, Inside the shaded and un-shaded greenhouses at 2.0 m from ground.
2
3 Fig.6 Solar radiation inside two parts of the shaded greenhouse, and un-shaded greenhouse compared to the outside
4 radiation.
5 The PE cover reflected more direct-beam radiation during afternoon. Consequently, it was found
6 that solar radiation under the PE cover with the partial shading of the BIPV on the greenhouse
7 was 30 %-35 % at 12:00 noon whereas, under the plastic cover it was only 75 %-80 % compared
8 to the direct sunlight outside the greenhouse on sunny days at a tilt angle of 30º. On the other

13
ACCEPTED MANUSCRIPT

1 hand, there was no observed variation of solar radiation on the cloudy days between the shaded
2 and un-shaded greenhouses. It can be also observed that the solar radiation in the shaded
3 greenhouse in the middle of the shaded area (shaded1) was lower than that of the east shaded
4 area (shaded 2) by an average of 100 W/m2 due to the movement of the shading area away from
5 it, and the effect of the unshaded half of the greenhouse (Fig 7a). The peak values of solar
6 radiation on clear and sunny days during the afternoon were high outside the greenhouse ranging
7 from 0.5-1.15 kW/m2 while, inside the greenhouse under the PE cover (Control) it ranged from
8 0.4-0.700 kW/m 2 and low under the PE with the BIPV ranged from 0.10-0.56 kW/m 2 as shown
9 in Fig.7b, these results are in agreement with Cossu et al. [20]. The greenhouse cover can absorb
10 and reflect the solar radiation. Accordingly, attaching PV panels inside greenhouses can decrease
11 the generated electric energy [9, 35]. Cossu et al.[8] reported that the shading of opaque PV
12 modules reduced the availability of solar radiation inside the greenhouse by 64 % up to 82 % for
13 the areas under the non-transparent PV covers. Meanwhile, they reported that the income derived
14 from the energy production was considerably higher than that resulting from the tomato crops. It
15 was reported that using the flexible PV and thin films, the semi-transparent PV panels, and the
16 spherical micro-cells, can increase the amount of solar light entering the greenhouse [32, 36, 37].
17 Accordingly, the BIPV can be considered as a moderate technology between the opaque PV and
18 the plastic cover, due to the light transmission of the semi-transparent PV being higher than that
19 of the non-transparent PV and lower than the single plastic cover and it can fit some of
20 greenhouse crops. Results revealed that when the average light intensity outside the greenhouses
21 was 86.3 klx (on a sunny day of November at 1 m from the ground), the average light intensity
22 under the plastic cover was 64 klx. Meanwhile, light intensity under the plastic cover with the
23 BIPV was 27 klx. These results revealed that the plastic cover transmittance was about 75 % and
24 the transmittance of the plastic cover with the BIPV was about 30 % compared to the outside
25 light intensity. Thereby, the combination of the BIPV and plastic cover decreases the light
26 intensity inside the greenhouse by approximately 35 %- 40 % compared to the single plastic
27 cover.

14
ACCEPTED MANUSCRIPT

1600
Un-shaded Outside Shaded 1 Shaded 2
1400

1200
Solar radiation (W/ m2)

1000

800

600

400

200

0
7:01
8:03
9:05
10:01
11:03
12:50
13:46
14:48
15:50
16:46 7:13
8:15
9:16
10:13
11:15
12:16
13:13
14:15
15:16
16:13
17:15
18:16

Time (h)

(a) 16th -17th June, 2017 (cloudy) at a1.0 m from ground


1

Outside Shaded Un-shaded


1400

1200
Solar radiation (W/m2)

1000

800

600

400

200

0
0:00 6:00 12:00 18:05 0:05 6:05 12:05 18:05 5:44 11:44 17:44 23:44

Time (h)
2
(b) 15th -17th Febraury, 2016 at 2.0 m from ground
3 Fig.7 Average solar radiation, inside the shaded greenhouse with the BIPV, un-shaded greenhouse and outside the
4 greenhouse.
5
6

15
ACCEPTED MANUSCRIPT

1 The amount of light in the wavelength range of 400 to 700 nm can be defined as
2 Photosynthetically Active Radiation (PAR) and it can be measured in either micromoles of
3 photons per square meter per second (µmol/ m2/s) or in watts per square meter (W/m2). Results
4 showed that the solar radiation under the BIPV shading in the peak period during daytime was
5 higher than or equal to 200 W/m2 from 10:00 a.m. to 5:00 p.m. The minimum amount of
6 irradiation necessary to ensure sufficient growth and flowering corresponds to a daily global
7 radiation of 2.0 -2.3 kWh/m2 days, (200 W/m2 equal to 1000 µmol/ m2/s).
8 Consequently, the small reduction in light transmitted through the PE cover compared to the
9 combination of the PE with the BIPV did not affect the growth of the tomato plants.
10 The position and orientation of PV modules on greenhouse roofs must be considered carefully to
11 provide a sufficient electrical energy with minimum shading [9-11]. Thereby, for an east-west
12 oriented greenhouse, the south-facing roof is suitable for generating electricity by PV modules,
13 but the effects of shading by the PV modules will be great. Moreover, it was observed that lots of
14 dust accumulates on the plastic cover of the greenhouse and under the BIPV, this dust could
15 prevent the light from entering the greenhouse. Thus, a regular cleaning system is recommended
16 to clean the dust on the BIPV and the plastic cover.
17 3.3. Electric energy production of BIPV on greenhouse
18 Figure 8 illustrates that the average of the generated electric energy of BIPV increased gradually
19 with the increasing of solar radiation on sunny days. The generated electric power at 10:00 a.m.
20 was 240 W and reached the peak value at 1:00 p.m. with 414 W eventually decreasing again to
21 235 W at 5:00 p.m.

22

16
ACCEPTED MANUSCRIPT

Generated power (Wp) 500

400

300

200

100

0
9:59 11:00 12:00 13:00 13:59 15:00 16:00 16:59 18:00
Time (h)
2

3 Fig. 8 Average of the generated electric energy from BIPV on 14th February, 2016.

4 The daily total greenhouse power consumption of ventilation and lighting systems in a typical
5 day of July was 1.85 kWh/ day. According to the meteorological data as shown earlier in Table 3
6 and the reported data that the optimal temperature range for most plants is between 22 and 28 ºC
7 in the daytime and 15-20 ºC at night [38], the annual required energy for heating greenhouses in
8 Kunming is only required in December, January and February. While, the cooling is only
9 required in May, June and July along with natural ventilation and shading as well as no need for
10 lighting. Accordingly, this integrated system can provide most of the annual energy demand for
11 heating and cooling systems. The semi-transparent BIPV could produce enough electricity to
12 meet the power consumption during the period from March through November. Nevertheless, the
13 required energy for heating demand will be higher than the generated electric energy from the
14 BIPV from December to February.
15 Results of the PolySun simulation showed that the best tilt angle to generate more electric energy
16 of the BIPV when it faced the South direction was 25º and the best tilt angle when it faced the
17 North direction was 20º. Meanwhile, the CO2 savings at those angles was higher than the other
18 angles. The annual generated electric energy of the BIPV facing south at 25º and roof area of 20 %
19 was 638 kWh and the CO2 savings were 342.1 kg whereas, at 20º and roof area of 20 % facing
20 north was 493 kWh and the CO2 savings were 264.5 kg. It was reported that the specific CO2
21 emission of the rooftop grid-connected PV systems was in the range of 50-100 g/kWh generated

17
ACCEPTED MANUSCRIPT

1 while the CO2 emissions of burning fossil fuels could be 504-900g/ kWh generated [39, 40].
2 Therefore, the BIPV has a great impact on the reduction of CO2 emissions.

3 On the other hand, by increasing the occupied south roof area to be 40 % at tilt angle of 25º it
4 produced 1247 kWh while, at the north roof area of 40 % and tilt angle of 20º it produced 968
5 kWh as shown in Tables 4 and 5.

6 Table 4. The annual generated electric energy (kWh) of the PV modules at different tilted angles
7 and directions at 20 % of the greenhouse roof area

Tilted 20º 25º 30º 35º 40º 45º 50º 55º


Facing South 635 638 637 632 624 612 597 579
CO2 savings( kg) 340.8 342.1 341.5 339 334.6 328.4 320.5 310.8
Facing North 493 462 430 396 362 332 306 282
CO2 saving (kg) 264.5 248 230.4 212.3 194.3 178.2 164.4 151.5
8
9 Table 5. The annual generated electric energy (kWh) of the PV modules at different tilted angles
10 and directions at 40 % of the greenhouse roof area.
11

Tilt angle 20º 25 º 30 º 35 º 40 º 45 º 50 º 55 º


Facing South(kWh) 1242 1247 1244 1235 1219 1197 1168 1133
CO2 savings (kg) 666.3 668.6 667.4 662.4 653.9 641.9 626.5 607.7
Facing North(kWh) 968 908 844 778 713 654 604 557
CO2 savings 519.1 486.9 452.8 417.5 382.3 351 324 298.9
12

13 The generated electric energy of the system facing the South direction from May to October was
14 less than that from November to April. This is may be due to the lower attitude angle of the sun
15 in this period, which can cause more shading on the panels and resulted in varies irradiation
16 levels on the surface of BIPV modules. Therefore, the monthly peak production of electric
17 energy from the south roof at covering area of 20 % and tilt angle of 30º was reached in March,
18 with 68 kWh and the lowest was found in July at 41 kWh. In contrast, the generated electric
19 energy of the system facing the north direction from March to September was higher than that
20 from November to February. Thereby, the peak production of electric energy from the North roof
21 at tilt angle of 20º was reached in April with 59 kWh and the lowest was found in December at
22 25 kWh as shown in Fig.9.

23

18
ACCEPTED MANUSCRIPT

140
a
Energy output (kWh/month)
South 40%
120
b North 40%
120
100
100
80
80

60
60

40 40
Jan Feb Mar Apr May Jun Jul Aug Sept Oct Nov Dec Jan Feb Mar Apr May Jun Jul Aug Sept Oct Nov Dec
Energy output (kWh/month)

70 C 60 d
South 20% North 20%
60
50
50
40
40

30
30

20 20
Jan Feb Mar Apr May Jun Jul Aug Sept Oct Nov Dec Jan Feb Mar Apr May Jun Jul Aug Sept Oct Nov Dec

Month Month
1

2 Fig. 9 Average of the monthly generated electric energy at different roof area and tilt angles. (a)Roof area 40 %
3 facing South at tilt angle of 25ᵒ, (b) Roof area 40 % facing North at tilt angle of 20ᵒ, (c) Roof area 20 % facing South
4 at tilt angle of 25ᵒ, (d) Roof area 20 % facing North at tilt angle of 20ᵒ.

5 It was reported that the annual electric energy consumption per unit greenhouse area in different
6 countries was deviated from 0.1 - 528 kWh/m2 for heating, cooling, lighting, and irrigation
7 systems according to the location and the environmental control technology in the greenhouse
8 [16]. Thus, the results of this experiment will be applicable for many regions around the world.

9 In this study a semi-transparent mono-crystalline silicon double glazed PV panels were used for
10 a high transmittances, which was at the same price of the opaque PV modules. Thus, using a high
11 efficiency BIPV can be recommended and the extra income from the sale of the remaining
12 energy to the grid could recover the extra costs after a period of time.

13

14

19
ACCEPTED MANUSCRIPT

1 3.4 Effect of BIPV shading on tomato

2 Normally, different plants require different amounts of light for optimal growth. The movable
3 shading curtains can reduce the energy load in the greenhouse crop during warm and sunny
4 conditions and they reduce heat radiation losses at night [41]. Shading not only affects plant
5 morphology, physiology and microstructure but also has an important impact on the total
6 biomass production. Plant growth requires appropriate photosynthetic photon flux density
7 (PPFD); excessively high or low intensity will prevent photosynthesis in the plant [42, 43].
8 Mashonjowa et al.[44] reported that shading a greenhouse by whitening reduced the transmission
9 coefficients for photosynthetically active radiation, total solar and thermal radiation of the
10 greenhouse cover from 0.75 to 0.53; 0.74 to 0.55 and 0.45 to 0.43, respectively. The plant height
11 and number of leaves under BIPV shading were insignificantly higher than those of the un-
12 shaded plants. In addition, it can be clearly seen that the chlorophyll contents of leaves in
13 shading were higher than those of the control. However, the differences were not significant in
14 chlorophyll contents between plants in control and shading greenhouses. The relative chlorophyll
15 content in leaves estimated by the SPAD-502 chlorophyll meter can be an efficient way to
16 evaluate plant nitrogen (N) status in many crops and some tree species [45]. Therefore, it can be
17 confirmed that there were no differences in the chlorophyll contents and nitrogen concentration
18 in tomato leaves under BIPV shading and the control as shown in Fig. 10. It was observed that
19 the air temperatures and solar radiation in the un-shaded greenhouse were higher than that in the
20 shaded greenhouse with the BIPV. Consequently, the tomato pots were dried quickly and more
21 water was consumed because the tomato transpiration rate increased with the increasing of air
22 temperature. These results are consistent with the previous studies of [46]. On the other hand,
23 tomato plants grown in hot climates for the entire season under a shade of 30 % to 40 % had a
24 higher yields and more fruit than those grown without shade [47]. It was found that the light
25 intensity of 300 μmol/ m2/s was more suitable for the culture of tomato seedlings in terms of the
26 fresh weight, dry weight, stem diameter and health index [48]. Consequently, the integration of
27 BIPV and greenhouses could be more applicable in hot climates to grow tomatoes. Similar
28 results were reported by Raual et al.[15].
29

20
ACCEPTED MANUSCRIPT

Control Treatment
120
Plant height (cm)
100
80
60
40
20
0
37 days 47 days 67 days 87 days
Number of leaves

30

20

10

0
37 days 47 days 67 days 87 days
Chlorophyll content (mg/g)

60

50

40

30

20

10

0
37 days 47 days 87 days

Days after treatment (DAT)

1 Fig. 10 Plant height, number of leaves chlorophyll content of tomato. Vertical bars represent mean ± standard errors
2 at P <0.05, (n = 20 and n=50 for chlorophyll samples).
3 In addition, the leaf area of shading plants under the BIPV was insignificantly bigger than those
4 leaf areas without shading after 37 days and 47 days of shading, respectively (Fig.11).

21
ACCEPTED MANUSCRIPT

Leaf area (cm2)


350 Control Treatment
300
250
200
150
100
50
0
37 days 47days

Days after treatment (DAT)


1

2 Fig. 11 leaves area. Vertical bars represent mean ± standard errors at P <0.05, (n = 20).

3 3.5 Economic analysis

4 The photovoltaic systems can be evaluated by the economic and environmental issues when the
5 genereated energy used for internal electric consumption or for sale to the grid [49].The payback
6 period can be calculated by dividing the amount of the initial investment by the cumulative net
7 cash flow for each period. The total costs of the photovoltaic systems include the cost of the
8 structure, PV panels, inverters, handling system, and storage batteries as well as the maintenance
9 and insurance costs (1 %) of the initial cost [50]. In the present study, the BIPV panels, the
10 structure, micro- inverters, wirings and the system maintenance and insurance costs were 5000
11 RMB ($720) and the annual electric energy production of the BIPV system at 20 % of the
12 greenhouse roof area, was 637 kWh (24.5 kWh/ m2 ) of greenhouse floor area). The price of 1
13 kWh in China is about 0.50 RMB [51]. Thus, the annual income of BIPV system will be 319
14 RMB. This income could be increased by increasing the BIPV covering area and efficiency. The
15 reduction of the annual energy production was set at 1 %, due to the reduction of the BIPV
16 efficiency, the lifespan of the BIPV panels is assumed to be 25 years and the rate of increase in
17 energy prices was assumed to be 3 % as well as the inflation of 2 % and interest of 3 %.
18 Consequently, the payback period of using the BIPV system was nine years and this period of
19 time could be shortened by choosing the most suitable plants to increase the annual yield
20 production. On top of that, the environmental benefits of using the BIPV panels are considerable,
21 due to the reduction of CO2 emission. Therefore, the integration of semi-transparent PV with
22 greenhouses would be an appropriate system in subtropical regions. Nevertheless, the PV panel’s

22
ACCEPTED MANUSCRIPT

1 technologies have been specifically developed for greenhouse applications; further investigation
2 for testing the performance on the field and the impact on the crops still needed. It was indicated
3 in 2012 that the payback period to return the investment capital of integrated PV panels on
4 greenhouses would be about 18 years in Spain [15]. While, in 2016 Marucci and Cappuccini [52]
5 reported that the calculated payback period of a dynamic photovoltaic greenhouse was 6 years in
6 clear sky conditions in Italy. Subsequently, the combination of BIPV and greenhouses could be a
7 beneficial option for the remote areas in the near future due to the reduction of PV prices and
8 governmental support.
9 4. Conclusion
10 The integration of semi-transparent BIPV panels with a transmittance ratio of 47 % on the roof
11 of greenhouses not only decreases the energy load, but also generates appropriate energy for the
12 supplemental lighting and heating in winter as well as forced ventilation in summer. The annual
13 generated electric energy of the BIPV panels per unit floor area of greenhouse was ranged from
14 24.5 kWh/m2 to 47.5 kWh/m2 at 20 % to 40 % of greenhouse roof covering area. Furthermore,
15 the proper tilt angle for generating more electric energy of the BIPV on the South roof of
16 greenhouse was 25º and the best tilt angle in the North roof was 20º. The BIPV shading
17 decreases the air temperature by approximately (1 ºC-3 ºC) on clear days and insignificantly
18 increases the relative humidity by 2 % as well as decreases the solar radiation by 35 % - 40 %
19 compared to un-shaded greenhouse. Moreover, The BIPV insignificantly affected the growth of
20 tomato. Results also revealed that the payback period of using the BIPV system was 9 years for
21 only 20 % of the greenhouse roof area. The annual electric energy production of The BIPV can
22 be increased particularly in high-irradiation regions and remote areas. Therefore, it can be
23 recommended that the most suitable crops, vegetables and ornamental plants under the BIPV
24 system those which need low light intensities. Furthermore, an external integration of a mobile
25 BIPV with a high efficiency or spherical micro-cells as a semi-transparent photovoltaic (PV)
26 technology on the greenhouse roof are highly recommended in the subtropical regions. In
27 addition, further studies should be conducted to explore the optimal integration of a new PV
28 panels with different plants in agricultural greenhouses.

29

30

23
ACCEPTED MANUSCRIPT

1 Acknowledgements

2 The authors acknowledge the financial support of the international scientific and technological
3 cooperation projects of China (No. 2011DFA62380) and the collaborative innovation center of
4 research and development of renewable energy in the southwest area of China (No.
5 05300205020516009)

6 References

7 [1] B.A. Keating, M. Herrero, P.S. Carberry, J. Gardner, M.B. Cole, Food wedges: Framing the global food
8 demand and supply challenge towards 2050, Global Food Security, 3 (2014) 125-132.
9 [2] J. Bundschuh, G. Chen, T. Yusaf, S. Chen, J. Yan, Sustainable energy and climate protection solutions
10 in agriculture, Appl Energ, 114 (2014) 735-736.
11 [3] M. Goe, G. Gaustad, Strengthening the case for recycling photovoltaics: An energy payback analysis,
12 Appl Energ, 120 (2014) 41-48.
13 [4] M. Djevic, A. Dimitrijevic, Energy consumption for different greenhouse constructions, Energy, 34
14 (2009) 1325-1331.
15 [5] R. Pahlavan, M. Omid, A. Akram, Energy use efficiency in greenhouse tomato production in Iran,
16 Energy, 36 (2011) 6714-6719.
17 [6] N. Kuswardhani, P. Soni, G.P. Shivakoti, Comparative energy input–output and financial analyses of
18 greenhouse and open field vegetables production in West Java, Indonesia, Energy, 53 (2013) 83-92.
19 [7] M. Canakci, I. Akinci, Energy use pattern analyses of greenhouse vegetable production, Energy, 31
20 (2006) 1243-1256.
21 [8] M. Cossu, L. Murgia, L. Ledda, P.A. Deligios, A. Sirigu, F. Chessa, A. Pazzona, Solar radiation
22 distribution inside a greenhouse with south-oriented photovoltaic roofs and effects on crop productivity,
23 Appl Energ, 133 (2014) 89-100.
24 [9] A. Yano, A. Furue, M. Kadowaki, T. Tanaka, E. Hiraki, M. Miyamoto, F. Ishizu, S. Noda, Electrical
25 energy generated by photovoltaic modules mounted inside the roof of a north–south oriented
26 greenhouse, Biosyst Eng, 103 (2009) 228-238.
27 [10] M. Kadowaki, A. Yano, F. Ishizu, T. Tanaka, S. Noda, Effects of greenhouse photovoltaic array
28 shading on Welsh onion growth, Biosyst Eng, 111 (2012) 290-297.
29 [11] A. Yano, M. Kadowaki, A. Furue, N. Tamaki, T. Tanaka, E. Hiraki, Y. Kato, F. Ishizu, S. Noda, Shading
30 and electrical features of a photovoltaic array mounted inside the roof of an east–west oriented
31 greenhouse, Biosyst Eng, 106 (2010) 367-377.
32 [12] H. Marrou, L. Dufour, J. Wery, How does a shelter of solar panels influence water flows in a soil–
33 crop system?, Eur J Agron, 50 (2013) 38-51.
34 [13] R.H.E. Hassanien, M. Li, Influences of greenhouse-integrated semi-transparent photovoltaics on
35 microclimate and lettuce growth, Int J Agric & Biol Eng, 10 (2017) 11-22.
36 [14] C. Dupraz, H. Marrou, G. Talbot, L. Dufour, A. Nogier, Y. Ferard, Combining solar photovoltaic panels
37 and food crops for optimising land use: Towards new agrivoltaic schemes, Renew Energ, 36 (2011) 2725-
38 2732.
39 [15] U.-S. Raúl, Ángel Jesús Callejón Ferre, José Pérez Alonso, Ángel Carreño Ortega, Greenhouse tomato
40 production with electricity generation by roof-mounted flexible solar panels, Sci Agr, 69 (2012) 233-239.
41 [16] R.H.E. Hassanien, M. Li, W. Dong Lin, Advanced applications of solar energy in agricultural
42 greenhouses, Renew Sustain Energy Rev., 54 (2016) 989-1001.

24
ACCEPTED MANUSCRIPT

1 [17] C. Breyer, O. Koskinen, P. Blechinger, Profitable climate change mitigation: The case of greenhouse
2 gas emission reduction benefits enabled by solar photovoltaic systems, Renew Sustain Energy Rev, 49
3 (2015) 610-628.
4 [18] R. Foster, A. Cota, Solar water pumping advances and comparative economics, Energy Procedia, 57
5 (2014) 1431-1436.
6 [19] A.J. Veldhuis, A.H.M.E. Reinders, Reviewing the potential and cost-effectiveness of off-grid PV
7 systems in Indonesia on a provincial level, Renew Sustain Energy Rev, 52 (2015) 757-769.
8 [20] M. Cossu, L. Ledda, G. Urracci, A. Sirigu, A. Cossu, L. Murgia, A. Pazzona, A. Yano, An algorithm for
9 the calculation of the light distribution in photovoltaic greenhouses, Solar Energy, 141 (2017) 38-48.
10 [21] H. Fatnassi, C. Poncet, M.M. Bazzano, R. Brun, N. Bertin, A numerical simulation of the photovoltaic
11 greenhouse microclimate, Solar Energy, 120 (2015) 575-584.
12 [22] B. Norton, P.C. Eames, T.K. Mallick, M.J. Huang, S.J. McCormack, J.D. Mondol, Y.G. Yohanis,
13 Enhancing the performance of building integrated photovoltaics, Sol Energy, 85 (2011) 1629-1664.
14 [23] D.H.W. Li, T.N.T. Lam, W.W.H. Chan, A.H.L. Mak, Energy and cost analysis of semi-transparent
15 photovoltaic in office buildings, Appl Energy, 86 (2009) 722-729.
16 [24] G.N. Tiwari, H. Saini, A. Tiwari, A. Deo, N. Gupta, P.S. Saini, Periodic theory of building integrated
17 photovoltaic thermal (BiPVT) system, Sol Energy, 125 (2016) 373-380.
18 [25] H. Manz, U.-P. Menti, Energy performance of glazings in European climates, Renw Energy, 37 (2012)
19 226-232.
20 [26] Y. Wang, R.H.E. Hassanien, M. Li, G. Xu, X. Ji, An experimental study of building thermal
21 environment in building integrated photovoltaic (BIPV) installation Bulgarian Chemical Communications,
22 48 (2016) 158 - 164.
23 [27] H.A. Ahemd, A.A. Al-Faraj, A.M. Abdel-Ghany, Shading greenhouses to improve the microclimate,
24 energy and water saving in hot regions: A review, Sci Hortic-Amsterdam, 201 (2016) 36-45.
25 [28] C. Li, H. Wang, H. Miao, B. Ye, The economic and social performance of integrated photovoltaic and
26 agricultural greenhouses systems: Case study in China, Appl Energy, 190 (2017) 204-212.
27 [29] P.J. Sonneveld, G.L.A.M. Swinkels, J. Campen, B.A.J. van Tuijl, H.J.J. Janssen, G.P.A. Bot, Performance
28 results of a solar greenhouse combining electrical and thermal energy production, Biosy Eng, 106 (2010)
29 48-57.
30 [30] F. Al-Shamiry, Ahmad D, Sharif ARM, Aris I, Janius R, Kamaruddin R, Design and development of
31 photovoltaic power system for tropical greenhouse cooling, Am J Appl Sci 4(2007) 386–389.
32 [31] A. Al-Ibrahim, Al-Abbadi N, Al-Helal I., PV greenhouse system, system description, performance and
33 lesson learned, Acta Hortic (ISHS) 710 (2006) 251-264.
34 [32] A. Yano, M. Onoe, J. Nakata, Prototype semi-transparent photovoltaic modules for greenhouse roof
35 applications, Biosyst Eng, 122 (2014) 62-73.
36 [33] G. Trypanagnostopoulos, A. Kavga, Μ. Souliotis, Y. Tripanagnostopoulos, Greenhouse performance
37 results for roof installed photovoltaics, Renew Energy, 111 (2017) 724-731.
38 [34] J. Paul W. Stackhouse, NASA Surface meteorology and Solar Energy, in, 2015
39 [35] J.G. Pieters, J.M. Deltour, M.J. Debruyckere, Light transmission through condensation on glass and
40 polyethylene, Agr forest meteorol, 85 (1997) 51-62.
41 [36] M. Cossu, A. Yano, Z. Li, M. Onoe, H. Nakamura, T. Matsumoto, J. Nakata, Advances on the semi-
42 transparent modules based on micro solar cells: First integration in a greenhouse system, Appl Energy,
43 162 (2016) 1042-1051.
44 [37] A. Marucci, D. Monarca, M. Cecchini, A. Colantoni, A. Manzo, A. Cappuccini, The Semitransparent
45 Photovoltaic Films for Mediterranean Greenhouse: A New Sustainable Technology, Math Probl Eng,
46 2012 (2012) 14.
47 [38] Castilla N, J. Hernandez, Greenhouse technological packages for high quality production, Acta Hortic,
48 761 (2007) 285-297.

25
ACCEPTED MANUSCRIPT

1 [39] E.A. Alsema, Energy pay-back time and CO2 emissions of PV systems, Progress in Photovoltaics:
2 Research and Applications, 8 (2000) 17-25.
3 [40] M. Pehnt, Dynamic life cycle assessment (LCA) of renewable energy technologies, Renew Energy, 31
4 (2006) 55-71.
5 [41] I.M. Al-Helal, A.M. Abdel-Ghany, Responses of plastic shading nets to global and diffuse PAR
6 transfer: Optical properties and evaluation, NJAS - Wageningen J Life Sci, 57 (2010) 125-132.
7 [42] Y. Dai, Z. Shen, Y. Liu, L. Wang, D. Hannaway, H. Lu, Effects of shade treatments on the
8 photosynthetic capacity, chlorophyll fluorescence, and chlorophyll content of tetrastigma hemsleyanum
9 diels et gilg, Environ Exp Bot, 65 (2009) 177-182.
10 [43] D. Zhao, Z. Hao, J. Tao, Effects of shade on plant growth and flower quality in the herbaceous peony
11 (Paeonia lactiflora Pall.), Plant Physiol Bioch, 61 (2012) 187-196.
12 [44] E. Mashonjowa, F. Ronsse, T. Mhizha, J.R. Milford, R. Lemeur, J.G. Pieters, The effects of whitening
13 and dust accumulation on the microclimate and canopy behaviour of rose plants (Rosa hybrida) in a
14 greenhouse in Zimbabwe, Sol Energy, 84 (2010) 10-23.
15 [45] M.C. Bonneville, J.W. Fyles, Assessing variations in SPAD‐502 chlorophyll meter measurements
16 and their relationships with nutrient content of trembling aspen foliage, Commun Soil Sci Plan, 37 (2006)
17 525-539.
18 [46] M.P.N. Gent, Density and duration of shade affect water and nutrient use in greenhouse tomato., J
19 Am Soc Hortic Sci, 133 (2008) 619-627.
20 [47] A.M.R. Abdel-Mawgoud, S.O. El-Abd, S.M. Singer, A.F. Abou-Hadid, T.C. Hsiao, Effect of shade on
21 the growth and yield of tomato plants, in, International Society for Horticultural Science (ISHS), Leuven,
22 Belgium, 1996, pp. 313-320.
23 [48] X.-X. Fan, Z.-G. Xu, X.-Y. Liu, C.-M. Tang, L.-W. Wang, X.-l. Han, Effects of light intensity on the
24 growth and leaf development of young tomato plants grown under a combination of red and blue light,
25 Scientia Horticulturae, 153 (2013) 50-55.
26 [49] A. Fernández-Infantes, J. Contreras, J.L. Bernal-Agustín, Design of grid connected PV systems
27 considering electrical, economical and environmental aspects: A practical case, Renw Energy, 31 (2006)
28 2042-2062.
29 [50] A.N. Celik, Present status of photovoltaic energy in Turkey and life cycle techno-economic analysis
30 of a grid-connected photovoltaic-house, Renew Sustain Energy Rev, 10 (2006) 370-387.
31 [51] Y. Qiu, M. Li, R.H.E. Hassanien, Y. Wang, X. Luo, Q. Yu, Performance and operation mode analysis of
32 a heat recovery and thermal storage solar-assisted heat pump drying system, Solar Energy, 137 (2016)
33 225-235.
34 [52] A. Marucci, A. Cappuccini, Dynamic photovoltaic greenhouse: Energy efficiency in clear sky
35 conditions, Appl Energy, 170 (2016) 362-376.

36
37

26
ACCEPTED MANUSCRIPT

Highlights

1. A specific integration of a semi-transparent photovoltaic (BIPV) with


greenhouse was studied.
2. The generated electric energy of BIPV and the payback period were
estimated.
3. The ideal tilt angle to generate electric energy at the south roof was at 25º
and at the north roof was 20º.
4. The BIPV shading decreased the light intensity and air temperatures without
harming the tomato growth.

You might also like