You are on page 1of 20

0032-1

Lessons Learned from Labyrinth Type of Air Preconditioning


in Exergy-Aware Greenhouses
Birol Kilkis
Department of Aerospace Engineering
OSTIM Technical University, Ankara Turkey
e−mail: birolkilkis@hotmail.com

ABSTRACT
An exergy-based performance model shows that a labyrinth system may not save energy unless
the exergy of the electrical power demand due to the electrical fans and pumps involved in the
system is less than the thermal power exergy gained. Based on this model, exemplary exergy
and carbon-aware solar greenhouse concept was developed to decouple food products from
fossil fuel use. Sample design calculations for such a solar greenhouse in a region located north
of Zwolle in the Netherlands show that it may eliminate the use of fossil fuels, except grid
connections. New terms define nearly-zero exergy greenhouse and nearly-zero carbon
greenhouse. The paper further discusses that exergy-smart greenhouses with local farming
close the decarbonization loop when coupled with the urban life in slow cities with low-
temperature district energy systems.
KEYWORDS
Nearly zero-Exergy greenhouse, nearly zero-carbon greenhouse, heat pipe technology, exergy
destructions, Rational Exergy Management Model, Labyrinth air preheating and cooling, CO2
emissions responsibility, solar greenhouse
INTRODUCTION
Labyrinth type of underground pre-heating or pre-cooling of the ambient air intake in energy-
smart greenhouses, as well as green buildings, are getting broader attention with total
electrification and decarbonization strategies of the EU with 100% renewables [1]. However,
any such application, which seems to be highly energy efficient with a COP of much greater
than one, maybe responsible for nearly-avoidable CO2 emissions in proportion to the exergy
destructions due to the mismatch between the unit exergy of electrical fan power demand and
the thermal power exergy gained. Previous studies in the literature neglected this issue and
considered only the 1st Law performance. On the other side, using solar PV systems at an
increasing rate to reduce or eliminate the carbon footprint of greenhouses may not adequately
respond to this problem because all renewable energy systems carry similar emissions
responsibilities, even in the absence of fossil fuels. Solar PVT systems may be more responsive
to the problem if low-temperature heat demand is present. Nayak and Tiwari carried out
experimental and numerical analyses for the prediction of the performance of a photovoltaic-
thermal (PVT) collector integrated with a greenhouse in climatic conditions of Delhi, India.
Their exergy analysis of the PV-equipped greenhouse showed, however, quite a low value of
4% [2]. They extended their research to PV-operated labyrinth use for greenhouses for different
climatic conditions in India [3]. Another study revealed the significance of using heat pipes in
greenhouses, mainly due to the replacement of electrically operated pumps and fans, which
usually cause significant exergy destructions due to the difference between high unit exergy of
electrical power and the thermal power transmission in the greenhouse [4]. Ozgener, O., and
Ozgener, L. [5] presented a case study about an experimental greenhouse in which the fresh

1
0032-2

outdoor air is pre-heated in an underground labyrinth type of air-to-air heat exchanger system.
They carried out experiments during the shoulder season of October in the Aegean region. They
observed an average greenhouse heating capacity, QH of 7.67 kW, with an air temperature lift
of 6oC. Power demand, PE of the electric fan and motor is 0.736 kW. Its exergy demand, EXE
is 0.736 kW x 0.95 kW/kW ~ 0.7 kW. Here 0.95 kW/kW is the unit exergy of electric power.
This unit exergy brings the average COP value to 7.67 kW/0.7 kW, which is 10.96. The authors
reported the same COP as 10.51, most probably, by approximating the unit exergy of electric
power as one kW/kW. Although it is practical to round-off the unit exergy of electrical power
to one, it defies the ideal Carnot equation:
 Tref 
 = 1 −  (1-a)
 Tsup 

For any given reference environment temperature, Tref, the supply temperature Tsup must
approach infinity to render ε mathematically equal to one kW/kW. This temperature is not
practically possible in the observable universe [6]. On the other hand, in the solar system, for
example, the maximum radiant source temperature that the earth observes is 5778 K (Surface
temperature of the sun). Then for an average temperature of around 14.5oC on the earth's
surface layer, Equation 1 gives the ideal unit exergy for electricity to be 0.95 kW/kW.
According to the graphical data of Ozgener, O., and Ozgener, L. [5], the absolute fresh air inlet
to the labyrinth and the greenhouse inlet air temperatures were about, Tin=20oC (293 K) and
Tsup=26oC (299 K), respectively, giving a temperature lift of 6 K. According to Equation 1, the
thermal exergy gain of the labyrinth type of the earth heat exchanger, using their reported
experimental data is:
 T   293 K 
EXH = QH 1 − in  = 7.67  1-
 T   = 0.15 kW .
 sup   299 K 
On the other hand, as already mentioned above, the electrical power exergy is:

E XE = 0.736 kW×0.95 kW/kW= 0.7 kW ,


The comparison of these two results show that this system destroys exergy by an amount of the
difference between them:
EXdes = ( 0.7 kW-0.15kW ) = 0.55 kW .
This result concludes that this greenhouse heating application is responsible for exergy
destructions that lead to nearly-avoidable CO2 emissions responsibilities for externally
offsetting the exergy destruction. The above two exergy results give a simple exergy efficiency
of 21% (0.15 kW/0.7 kW), excluding parasitic exergy destructions. On the contrary, the
Authors reported an exergy efficiency of 89.25%, including exergy destruction at the ground
heat exchanger. This contradiction shows that a more comprehensive yet straightforward model
is needed to apply the 2nd Law in such applications, comprehensively and connected to the
emission responsibilities. In this token, and based on the same experimental greenhouse,
Hepbasli [7] reported that the exergy efficiency is 19.18%. This latter value is in good
agreement with the 21% exergy efficiency value given above.
To avoid adverse effects due to ingress of bacteria grown in the underground labyrinth wall
surfaces, dust, particulate matter, airborne viruses, and insects to the greenhouse or building
interiors [8] by directly allowing outdoor air from the labyrinth, a closed air loop separated by
a heat exchanger [9, 10], and or active filtering of the outdoor air, downstream the labyrinth
passage may be necessary. However, these measures increase the fan power required to
overcome the additional pressure losses at the heat exchanger and filter interfaces and
additional ducting. Therefore, a careful optimization between the air quality and power
consumption may be necessary. At any rate, the exergy of the power demand of fans and other

2
0032-3

ancillaries must be less than the thermal exergy gain from the underground labyrinth, as given
in Equation 1. Furthermore, the air quality may be indexed to the level of the particulate matter
in terms of PM2.5 and its equivalence to the CO2 emissions through the function (f) [11], which
allows optimizing the air quality versus power exergy demand. Equation 1-b is about the
balance between the CO2 equivalency of particulate ingress at an airflow rate of V (The first
term), CO2 emissions due to electricity supply from a central power plant (excludes nearly-
avoidable exergy destructions) to satisfy the additional power demand, PE, which is the
second term. The third term is the CO2 emissions savings indexed to an onsite natural gas boiler
with a CO2 content of 0.2 kg CO2/kW-h and an average boiler efficiency of 0.85. THE+F
corrects the third term for sensible heat according to the temperature drop taking place across
the heat exchanger and filters if they are present. This correction assumes that the flow rates
are the same. On an hourly basis, the result must be less than zero, indicating that the sum of
emissions savings is more than its CO2 emissions responsibility.
 0.2   T HE + F 
f ( PM 2.5) )V + PE cK PEF − QE    1 − 0 (1-b)
 0.85   Tin − Tsup 

In some cases, it might be possible to eliminate the primary fan power demand by introducing
natural air circulation through a sloped duct between the labyrinth beneath the greenhouse and
the greenhouse roof or between a multi-story building [12]. Rotta [13] investigated the seasonal
thermal storage capability of underground tunnels, where these tunnels may accumulate heat
in summer and utilize that heat in winter, thereby cooling the underground for the next cooling
season. Rotta identified seven design parameters. These are namely; 1- inlet, outlet, and mean
temperature of the heat transfer fluid (air), 2- thermal power storage or use in the labyrinth, 3-
thermal energy, 4 and 5- thermal losses and storage efficiency, 6- Specific storage capacity,
and 7-thermal stresses. However, this study, likewise most of the others cited below, did not
consider the exergy concept and did not include the 2nd Law analysis. Song, S., Song, J., and
Lim J. [14] have discussed the effectiveness of labyrinth systems in their paper concerning an
educational facility in South Korea. They experimented in an existing college building
equipped with a labyrinth system. They identified two cases, namely the outdoor air entering
directly to the air handler units on the ground level and outdoor air entering the air handlers
after passing through the labyrinth system. They have calculated the construction cost of the
labyrinth systems. After their seasonal measurements, they have concluded that peak loads
(including the latent loads) for the cooling and the heating seasons were reduced by 47.6% and
41.2%, respectively. A payback period of 12.1 years was estimated. Their calculations were
based on the 1st Law. With the COVID-19 Pandemic, 100% outdoor air, year-round, became
an essential part of the measures to be taken to minimize the virus spread. This necessity
increased the ventilation rates and fresh air intake, resulting in a significant increase in heating
and cooling loads, especially in large buildings like airport terminals, hospitals, and shopping
centers. Nevertheless, the downside of labyrinth systems, especially in hospital buildings, is
that outdoor air quality downstream of the labyrinth is a concern. The upside is that labyrinth
systems in a detached position may be retrofitted to existing buildings. Marmion, P. et al. [15]
have considered a basement type of labyrinth system for a conceptual 200-bed an acute-care
hospital with a water-based thermal mass instead of a conventional concrete-mass labyrinth.
Their main conclusion was that the labyrinth with a natural/hybrid ventilation system increased
the energy savings from 60% to 80% energy savings. Rim, Sung, and Kim [16] have compared
different alternatives for the labyrinth system with 412.68 m2 concrete underground space for
maximum heat transfer efficiency, defined in terms of COP. COP changed between 1.920 and
1.925, according to the length of the labyrinth walls, of which they considered three cases.
They further investigated the effect of the internal shape in four cases, which showed a COP
range between 2.0 and 2.5. They concluded that the highest COP values during the summer

3
0032-4

months due to larger temperature differences across the labyrinth. [17]. Misra et al. have
investigated the effect of the humidity of the soil on the energy and exergy efficiency of
underground tunnel type of heat exchangers. Their field studies have shown that the exergy
efficiency decreases sharply with the pipe length and stabilizes around 49% thereon after a pipe
length of 25 m. They have also determined that wet soil yields higher exergy efficiency [17].

THE NEED AND THE AIM OF THE PRESENT STUDY


Today, the annual energy consumption of the food industry in the EU countries is about 17%
of the total consumption [18]. Poorer countries spend more energy than EU countries, while
global average energy consumption also increases [19]. Although it seems a relatively small
percentage among other sectors, higher levels of water consumption and increasing global
temperature despite several 1st Law measures, agriculture is responsible for large amounts of
CO2 emissions, despite, statistics show that agriculture had only 3.2 % of final energy
consumption in the EU in 2018. The share of agriculture in final energy consumption was
highest in the Netherlands (8.1 %) [20]. Ever-increasing atmospheric anomalies further
enhance agricultural losses, thus increasing agricultural energy consumption to offset the
product and greenhouse losses, which tend to be unrecoverable with the current insurance
plans. Therefore, more resilient, energy-smart, more robust greenhouses are becoming a
necessity. Although energy systems are improving for greenhouses, the systems in use are not
exergy-rational, like labyrinth systems and ventilation units [21], in many cases, according to
the 2nd Law in terms of their responsibilities about the nearly-avoidable CO2 emissions.
Higher the exergy rationality, higher is the food quality due to energy and cost savings, which
may be allocated to investigate further R&D activities, better fertilizers, and technical
improvement. The Netherlands is one of the most successful countries in these respects.
Therefore, this study aims to develop exergy-based design criteria and hints for better
greenhouse performance that faces already `energy-smart` greenhouses like in the Netherlands.
The challenge taken in this paper is to develop `exergy-smart` greenhouses.
CONCEPT DESIGN
A new greenhouse design has been developed with the primary objective of harnessing solar
energy in the most efficient, and more importantly, in the highest rational exergy management
efficiency, ψR, which leads to minimum CO2 emission responsibility. For a greenhouse to be
green, ψR is required to be greater than 0.7, as was developed for green buildings and cities
[22]. Figure 1 shows the design. Another objective of this design against global warming is to
achieve nearly-zero carbon greenhouse (nZCGH) and nearly-zero exergy greenhouse
(nZEXGH) status. These definitions were adopted from earlier definitions, which were
developed for zero-buildings [23]:
nZCGH: Nearly-zero carbon building is a greenhouse, which has an annual EDR value of
equal to or greater than 0.80. EDR is based on the exergy of heat and power consumption.

(  CO ) / (QH + E ) X 
EDR = 1 −  {EDR > 0} (2)
2

0.63 kg CO 2 /kW-h

0.63 kg CO2/kW-h in the denominator is the reference value [23]. If there is an active CCS
(Carbon capture and storage) system in the greenhouse, this amount is subtracted from the
numerator after reducing it by the CO2 equivalent-power consumed for CCS (CCSnet). In this
case, EDR may be less than zero.

4
0032-5

(  CO ) / (QH + E ) X − CCSnet 
EDR = 1 −  {EDR > 1} (3)
2

0.63 kg CO 2 /kW-h

nZEXGH: Nearly- zero Exergy Building is an individual building or a building compound


connected to the district, which on an annual basis provides at least 80 % of the total exergy
(Maximum 20% exergy Deficit) of heat (Cold included) and power to the district or to itself as
the total exergy of heat and power received from the district or the demand of the individual
building. In an individual building, the exergy deficit must be offset by temperature supply and
demand adjustments, if possible. Otherwise, onsite exergy sources must add on. The
greenhouse operates on DC power, except using an inverter for the grid AC power converter
during emergencies only. Other primary features are short-listed below.

• Uneven span maximizes the PV surface area on the south-facing side of the roof. PV panels
are 60% transparent with an average power generation of efficiency of 15%. There are
several ongoing research and R&D activities to apply PV technology to the greenhouse
roofs, like the Organic Photovoltaics (OPV), without compromising the plant growth [24].
These PV panels may also help passively cooling the greenhouse during the summer in hot
climates by night radiation to the sky. Capillary type of heat pipes sandwiched to the
transparent PV panels serve both for night radiation by absorbing heat from the indoors
through the desiccant wheel and transferring it to the PV panels, radiating heat back to the
sky. On hot summer daytimes, PV cells need cooling. Then, the same heat pipes transfer
heat to the desiccant wheel to thermally charge for dehumidification, which often coincides
with the hot weather. PV cells supply power for the desiccant wheel. In cold climates, a
mechanical shutter system may isolate the greenhouse from cold outdoor temperatures to
reduce the heat loss from the roof. The top glass of the PV panels may be tempered type
with electric wires for ice and snowmelt in severe conditions.
• On the north side of the chromatic glass layer, automatically adjust the sunlight at pre-
determined intensities according to the solar insolation, brightness, plant growth, and
indoor thermal requirements for minimum thermal loads.
• In hot climates, PVT panels may be used instead of PV panels. Low-temperature heat may
be seasonally stored in a second TES unit with heat pipes for winter use in additional root-
zone heating and ice or snow melting off the immediate surroundings of the greenhouse
• Slow city-compatible design (See Figure 13): With global warming and increasing drought,
irrigation is getting more serious. In this case, hybrid wind power-and-heat turbines (See
Figure 14), coupled with this greenhouse concept and the onsite hydrogen economy, help
reduce and green the irrigation power demand.
• The system is an all-air, a closed cycle with air source heat pump and fans,
• Cut-off control for labyrinth operations saves from the fan exergy demand.
In this design, there are six distinct components and one set of options. These are:

1- Earth-to-Air Heat Exchanging (EAHE) Labyrinth (Figure 1),


2- Air Source Heat Pumps (ASHP), Two in Series Cascade (Figure 1),
3- PV Panels and or PVT (Figure 1),
4- Hybrid Wall Panel, with Heat Pipes and Porous Air Wall (Figures 2 and 3),
5- Root-Zone Heating with Buried Heat Pipes
6- Carbon Capture and Storage Unit, CCS
7- A Set of Options, Comprising micro-CHP, Biogas reactor, and TES

5
0032-6

Heat Pipe

Shutter

GREENHOUSE

Figure 1. General Layout of the all-DC, Exergy-Smart Green House System, nZEXGH;
nZCGH, ©2021 B. Kilkis, CCS: Carbon Capture and Storage

Figure 2. Sectional Top View of the Radiant Wall Panel ©2021B. Kilkis

Figure 3. Back-Side Cut-Away View of the Heat-Pipe Wall Panel (HWP) ©2021B. Kilkis

6
0032-7

DEVELOPMENT OF THE MODEL


This model is based on the Rational Exergy Management Efficiency, ψR, by identifying the
major exergy destructions [22], then predicting the CO2 emissions responsibility, and finally
calculating the EDR. For satisfying the objective and scope of this study, an earlier model being
developed for ERV systems in green buildings has been adopted for greenhouse labyrinth
systems with or without heat pumps [25]. The earlier model applies to the greenhouse labyrinth
type of heat recovery from the ground heat and the recirculated air.
Exergy Destructions and CO2 Emissions
According to Rational Exergy Management Model (REMM) in small temperature differences
at low temperatures, for a simple case, like PV panel (Figure 4):
 Tref 
1 − 
 des1  dem  TE 
 R = 1− = 1− {ψR > 0.7} (4)
 sup  sup  Tref 
1 − 
 T f 
For multiple (n) exergy destruction points in a complex system:
n

 desi
 R = 1− i =1
m
(5-a)

j =1
sup

Figure 4. Exergy Destructions in a PV Cell.

For example, in the greenhouse concept, there are two exergy supplies, namely the labyrinth
and the PV panels. Labyrinth provides thermal power, Q1 with unit exergy, εH of (1-To/T1), and
PV panels provide electrical power, E with unit exergy (0.95 kW/kW). Then from Equation
5-b, the value of Ʃεsup may be calculated:
Q 
0.95 +  H  1 
E
 sup =
Q 
(5-b)
1+  1 
E
 To 
 H = 1 −  (5-c)
 T1 

The labyrinth thermal power is subject to the following cut-off conditions, whichever comes
first, to avoid fan power consumption for too little thermal gain.

7
0032-8

Cut-off conditions:

To ≥ Tg in winter and To ≤ Tg in summer (i)

To − Tg  5 K (ii) (5-d)

According to Figure 4, PV the PV panel generates electrical power. TE is the panel frame and
approximately the cell temperature, which may be high in hot climates requiring active or
passive cooling to preserve the rated power generation efficiency. PV system shown in Figure
4 destroys the thermal exergy, which otherwise could be utilized as thermal exergy in useful
applications, like DHW. However, the coolant temperature must be low to preserve the power
generation efficiency. Therefore, if the thermal exergy is going to be utilized in an open system
like a shower, the temperature needs to be peaked to reduce the Legionella risk to humans. In
this case, more exergy is destroyed, which is represented by T`E in the figure. As a whole, the
PV cell generates power with high unit exergy (0.95 kW/kW) at a 1st Law efficiency of ηPV.
Therefore, it replaces CO2 emissions from the power stock by (-CO2). Here REX is the exergy-
based ratio of renewables in power generation. On the other hand, the same PV cell is
responsible for exergy destructions in an amount of CO2 because this amount of exergy
destruction must be offset in the thermal power stock by someone, somewhere, most likely by
using fossil fuels. Based on natural gas, the multiplier (c) is 0.27 kg CO2/kW-h. The multiplier
(cK) also has the same value if based on natural gas power plants. PEF is the primary energy
ratio. The current EU average is 2.5 (Fuel to Plug). The reference temperature, Tref is taken to
be 283 K in heating and 273 K in cooling in our studies. Tf is the Carnot cycle-equivalent solar
temperature given by Equation 6.
Tref (6)
T =
 I  0.95 
f
1−  n 
 1366 
Referring to Figure 1;

T1 T Tref T
 des1 = 1 − ;  des 2 = 1 − r ;  des 3 = 1 − ;  des 4 = 1 − a (7)
Tg Tm TE T2
Parasitic power losses due to chromatic glass, desiccant wheel, and in-the-ground heat pipes
are neglected. Fan exergy losses at F1, F2. and F3 are:

 0.95 T 
 desF 1 =  − 1− o  (8)
 Q1 / F1 T1 

 0.95 T 
 desF 2 =  − 1 − a  T2 >Tm {In Winter Heating} (9)
 Q 3 / F2 T2 

Q3 = P  COP = Q2 + P (10)

COP = d + e T2 − Tm (11)

 Tg 
 des 3 = 0.95 − 1 −  (12)
 TE 

8
0032-9

Ta (13)
 des 4 = 1 −
Tp

To
 des 5 = 1 − (14)
Ta

CO2 = 0.27    des (15)

CO2 = ( P + F1 + F2 + F3 ) X − ( I n APVPV )  ( 0.2PEF ) − CO2ccs − CO2 CHP (16)

CO2 CHP = QB  0.2  (1 − REX ) ( H /  BL ) + PEF E  (17)

SAMPLE CALCULATIONS

The new greenhouse design is parametrically investigated for the climatic conditions of the
city of Zwolle in the Netherlands. January is the coldest month with an average high
temperature of 4.6°C and an average low of 0.9°C (Figure 5). Hourly lowest temperature
decreases down to -8oC. Therefore, Zwolle has a moderately cold climate with low solar
intensity in winter (Figure 6), according to the following figures.

Figure 5. Monthly Average Outdoor Temperatures in Zwolle, the Netherlands [26]

Figure 6. Monthly-Average Sun Hours in Zwolle [27]


Figure 7 shows that the ground temperature close to the surface for labyrinth applications is
gradually increasing primarily due to the urban heat island effect (UHI) and climate change.
From this perspective, the slow city concept, which nears cities to agriculture, is essential. On
the other hand, the minor advantage of increasing the number of labyrinth applications near

9
0032-10

and in the cities, both in buildings and greenhouses with seasonal thermal storage, may reduce
the UHI and provide thermal energy. If only the climate change is considered for labyrinth
applications, the ground temperature close to the surface, Tg may be taken 11oC (284 K). With
the urban heat island effect (UHI), Tg may be assumed to be 2 K higher (286 K). To demonstrate
the model, three variables, namely, To (Representing Climate), Tg (Representing Urban
Connection and the Future Trends), and ((Representing Plants Grown), were parametrically
varied to investigate their effects on the greenhouse performance.

Figure 7. Variation of the Ground Temperature with Depth, Climate Change, and
Urbanization. The dotted line is for the period before 1900 [28]

1- Effect of Climate, To
The effect of the climate represented by the dry-bulb design outdoor temperature, To varies
between -10oC (263 K) and +5oC (278 K). The EAHE efficiency and Tg are assumed to be
constant. Only εdes1, and εdesF1 change with To. Figure 8 shows their variation for different Q1/E
ratios (See Equation 5-b) that also affect ψR. Therefore, the same graph may also be interpreted
for CO2. Figure 8 shows that supply exergy decreases with outdoor temperature increase
because it approaches the ground temperature, 285 K. A critical parameter is the (Q1/E) ratio,
which also changes with climatic conditions and the outdoor temperature. Supply exergy
decreases by increasing the (Q1/E) ratio because the unit exergy of power is higher than the
unit exergy of thermal power supplied by the labyrinth.
2- Greenhouse Indoor Temperature, Ta Required for Grown Products
Indoor design temperature is complicated and largely depends on the production. Some
vegetables, like tomatoes, require a two-zone temperature gradient. One of them is the root-
zone temperature. According to this concept design, root-zone heating is independently
accomplished by ground heat pipes. Root-zone heating also has natural convection and
radiation heat transfer components from the surface to the indoors and plants. Convection heat
transfer from the root zone surface helps to adjust the indoor air temperature. The air
temperature, however, is primarily controlled by the heat-pipe-controlled air-to-air forced
convection plus thermal radiation type of hybrid wall panels. These vertically positioned heat
pipes have a low profile like baseboard type of equipment.

10
0032-11

0,7

0,65

0,6

0,55
Σεsup

0,5 Q1/E =1

0,45 Q1/E =0.5


Q1/E =2
0,4

0,35

0,3
260 265 270 275 280
Outdoor Air Temperature, To, K

Figure 8. Variation of Σεsup with To and (Q1/E) ratio

Figure 9 shows only the variation of εdes4, assuming that the fan exergy destruction remains
constant in the limited operational range. T2 depends on the COP value of ASHP. Therefore,
Figure 9 shows solutions for different T2 values that represent different COP values during
operation. Higher the greenhouse indoor air temperature, exergy destruction decreases with
higher supply air temperatures from ASHP. However, any increase in the supply air
temperature for a given air temperature supplied by the EAHE, COP of the heat pump
decreases, and the power consumption of the heat pump increases.

0,14

0,12

0,1

0,08
εdes4

T2=308K
0,06
T2=303 K
0,04 T2=328 K

0,02

0
284 286 288 290 292 294 296
Ta, K

Figure 9. Variation of εdes4 with Ta and T2 (Heat Pump Supply Air Temperature)

RESULTS
1- Exergy-Based Limits on Heat Pump Technology
Heat pumps are also responsible for CO2. Therefore, special care must be paid for their
sustainability ranking. The new exergy-based heat pump performance metric, HPF to be
maximized in the heating mode, shows that the temperature range for exergy-rational operation
is relatively limited [25, 29].

11
0032-12

COP   dem  d + e T f − Tsup  {Maximize HPF ≥ 1} (18)


HPF = =  (1 − X HP )
0.95  0.95 
 
.  Tret  {0< XHP <1} (19)
 T  (
 dem = 1 − = 1 − X HP )
 sup 

For a given source temperature Tf, the lower limit of XHP, namely the (Tret/Tsup) ratio, is given
by Equation 18 for HPF =1. It is further subjected to Equation 20.
1 0.95
X HP  1 − {COP =d + e T f - Tsup } (20)
 d + e T f − Tsup 
 
 0.95 
 
CO2 = 0.63 ( 0.95 − 1 − X HP )  0.63 X HP − 0.0315 {Minimize} (21)

When XHP is one, then CO2 is maximum because no exergy is utilized. According to Equation
21, CO2 is zero if XHP is 0.05, which is not practical because the demand exergy is too small
for economically feasible heat pump use or the flow rate is too high, consuming excessive
electrical exergy. Simultaneous solution for minimum XHP and CO2 with simple
differentiation gives the maximum supply temperature at a given, Tf in Equation 22.
d 
Tsup max   + T f  T f {For minimum CO2} (22)
e 
Then for practical values of XHP, as shown in Figure 10, only a small feasible domain for
exergy-rational heat pump operations is left, namely XHP ≥ 0.88, which also limits HPF to a
maximum value of 1.5 at a supply temperature of 338 K, according to Equation 20. With these
limitations, CO2 emissions are largely unavoidable (See Equation 21). For example, if XHP is
0.88 CO2 is 0.58 kg CO2/kW-h. The corresponding minimum COP is 8.3, which is difficult
to achieve in practice unless the supply temperatures required are lowered accordingly by
considering the relationship between COP and the building insulation. Figure 11 shows similar
results between the supply temperatures of 330 K and 360 K. Results do not change when the
T value is made a function of the supply temperature (Smaller T with lower resource
temperature). For example, if T is 5 K, then CO2 is slightly more: 0.588 kg CO2/kW-h.
Although small T reduces equipment oversizing, the cost of additional pumping power in the
building must be considered. Therefore, local optimization is necessary to refine the
decarbonization strategies about the extensive use of heat pumps. At 5DE systems T must be
less than 10 K, which is a strong constraint to the design:
T 10 (23)
X HP = 1 −  1−
Tsup Tsup
Lessons Learned From Labyrinth Systems
Figure 12 shows a labyrinth system for outdoor air pre-heating or pre-cooling for a building
[26]. These systems are touted as energy-efficient and environmentally friendly systems with
high COP values. For this statement to be accurate, the electric power exergy input for the air-
circulating fan(s) in the labyrinth must be much less than the thermal power exergy
(supply/extraction) for a positive unit exergy gain. If, for example, the winter outdoor
temperature, tout is -15oC, and the pre-heated air supply temperature, tin is +2oC, then for each
kW of heat (1 kWH), the fan power must be less than 0.065 kWE.
273.15 − 15
1−
PFAN  273.15 + 2 0.062
 H = = < 0.065 kWE /kWH
1 kWH E 0.95 0.95

12
0032-13

Figure 10. Variation of HPF with Tsup. Tf = 310 K. Maximum Tsup is 345 K

Tout (24)
PFAN  1 − / 0.95
Tin

Figure 11. Variation of HPF with the Supply Temperature, Tsup Required from the Heat Pump

If for example, PFAN is 0.25 kWE per kWH of heat obtained from (Winter) or rejected to (in
Summer), then Equation 24 is not satisfied, and exergy destruction takes place:

EXdes   E  0.25 kWE −  H 1 kWH = 0.95×0.25-0.062 1= 0.175 kW


This destruction is responsible for nearly-avoidable CO2 emissions CO2.

CO2 = 0.27 des 2 + 0.63 des1  (1 − REX ) (25)


REX is the exergy-corrected percentage of the renewable energy mix in the stock. If, for
example, the total fan power is supplied by the roof-mounted PV panels, then REX is 1.
The labyrinth system is also responsible for direct CO2 emissions in a proportion of the fan
energy demand satisfied annually from the grid per unit preheating or precooling, R:

PEF
CO2 = ck (1 − REX ) R {Per kWE-h} (26)
COP

13
0032-14

Fan Fan
Electric Electric
Power Power

Figure 12. A Typical Labyrinth System Catalog That Does Not Mention the Fan Power
Required [30]. It Has Been Additively Illustrated in this Figure by the Author.

An example of a labyrinth type of application is a solar greenhouse [31]. In this application,


0.736 kW is the electric power demand of the fan for a precooling capacity of 4.36 kW.
This capacity corresponds to a COP of 5.9. However, the exergy-based COP, namely COPEX,
is less than one:
 298 K 
1- 
COPEX = COP  
313 K 
= 5.9  0.0504 = 0.297  1
0.95 kW/kW

Under summer design conditions, the outdoor air temperature is 40oC (313 K), and it is
precooled in the labyrinth to 25oC (298 K). According to Equation 24, the fan power demand
should not exceed the thermal power exergy, which is 0.0503 x 4.36 kW = 0.219 kW.
The installed fan capacity is more than three times this limit. Therefore, exergy destruction,
EXDES in terms of lost power exergy, takes place.

Exdes = 0.95 kW/kW x 0.736 kW - (1-298 K/313 K) kW/kW x 4.36 kW = 0.49 kW

CO2 = 0.63 kg CO 2 /kW-h×0.49 kW= +0.31 kg CO 2 /h {Power exergy is destroyed}

At the same time, this labyrinth saves 4.36 kW of cooling, which otherwise should be
accomplished by an electrically operated chiller with a COP of 3. Irrespective of whether the
electric power comes from the grid or an onsite PV system, these systems are responsible for
exergy destructions. For example, the exergy destruction of a PV system is explained in Figure
4. In this case, the net CO2 savings is calculated by considering the electric power demand of
the chiller to be satisfied by PV panels:

E = 4.36 kW/(COP=3) = 1.45 kW.

At this design capacity, the net CO2 savings with respect to a reference of grid power supply
by a natural-gas thermal power plant with ψR = 0.45 is:
CO2 = (1.45 kW×PEF×0.2 kg CO2 /kW-h )  0.45 = -0.725 kg CO2 /h ,
Alternatively, normalizing to 1 kW of cooling:

-0.725 kg CO2/h (1 kW/1.45 kW) = -0. 5 kg CO2/h,

Here, PEF is 2.5. 0.2 kg CO2/kW-h is the unit CO2 content of natural gas (cK). Therefore, the
labyrinth is responsible for 0.49 kg CO2 emissions in the sector per hour of operation at design

14
0032-15

conditions and saves only 0.50 kg CO2 emissions from the carbon stock. The net savings is
almost nil. This sample result shows that the unit exergy difference between electric power
(0.95 kW/kW) and quite a low unit thermal exergy of heat involved in a labyrinth is a crucial
design and rating factor to be considered both in theory and practice.
CONCLUSIONS
The agriculture and the food industry need to be coupled within cities and their surroundings
such that agriculture, animal farming, built environment, mobility, and other sectors involved
must be equipped with the same type of holistically conceptualized solutions to decouple global
warming with human needs and activities, by utilizing low-temperature sources. However,
coupling of widely available low-temperature renewable sources and waste heat requires new
exergy-based metrics. The greenhouse model presented in this paper may be coupled with low-
temperature district energy systems (5DE) with the same fundamental model leading to slow
cities that produce their food by renewable energy in advanced industrial type greenhouses,
which is also a growing need against weather anomalies. In this respect, the concept may be
expanded to green cities, low-exergy cities, and 100% renewable heating and cooling concepts
(100%RHC). The following two figures, namely Figure 13 and Figure 14, are examples in the
quest for such solutions.
1. Expanding the Greenhouse Concept
The Netherlands is a perfect country for the slow city concept because farming and cities,
except megacities, are already side-by-side and even mixed. Figure 13 shows such a holistic
option. The hydrogen economy is used both as an energy carrier and storage. Power generated
by the onsite wind turbines and PVT panels, either raised above the ground or attached to the
turbine tower, as seen in Figure 14, runs the irrigation process. In turn, irrigation water, which
is pumped from the water wells, cools PVT panels before irrigation. Rational Exergy
Management Efficiency, ψR increases above 0.70.
2. Hybridization of Renewables
As shown in Figure 14, wind and solar energy may be harvested on a single wind-turbine tower,
while agricultural land is spared instead of occupied by PV panels on the ground. PVT panels
are cooled in summer by embedded heat pipes connected to a hydronic collector in each panel.
Panels have a sun-tracking mechanism around the wind turbine tower. Heat may be seasonally
stored for the use of nearby farms and greenhouses. Surplus power is stored in the form of
pressurized air in the tower.
This paper has presented an advanced version of the recent trend about the Agri-voltaic system,
which is the integration of semitransparent organic solar cells into greenhouse construction.
This technology reduces heating and cooling loads by light filtration, which also helps to
control the photosynthesis action of the plants grown in the greenhouse [33]. The latter function
is supported on the narrow side of the slanted roof of the greenhouse concept introduced in this
paper, shown in Figure 1, where light control glass (Chromatic Glass) is blended with organic
PV cells. Agri-voltaic systems almost double the combined use of the land area, both
agriculture, and power generation [34]. However, the simple definition of Agri-voltaic system
has not yet been investigated from the 2nd Law of thermodynamics, which reveals the real
benefits of Agri-voltaic system, such as minimizing exergy destructions and incorporating
other green systems like root-zone heating from the ground by using heat pipes. Further
enhancements include humidity control by the desiccant wheel, which is driven by the power
and heat generated by the solar PVT system, micro-cogeneration by the onsite biogas
production from agricultural and animal waste (If available on site), thermal energy storage,
and biogas fuel storage, along with HWP (Heat-pipe wall panel) heaters in the greenhouse.

15
0032-16

Figure 13. From Green Houses to Food Industry and Slow Green Cities [32]
© 2020 B. Kilkis

Figure 14. Hybrid Wind Turbine-PVT system Concept for a Dutch Farm, © 2020 B. Kilkis

Heat-pipe-based design minimizes the use of electrically driven fans and pumps in the
greenhouse. Replacement of fans and pumps by heat pipes reduces parasitic losses. Air-source
heat pump (ASHP) contribution is minimized due to their ozone-depleting potential (ODP)
because of their refrigerant content. In order to avoid or minimize the ODP, CO2 captured from
the CCS system is used as the refrigerant. At the same time, multiple heat pumps are in cascade

16
0032-17

such that their overall COP approaches 10, which is the threshold for an exergy balance
between their power demand exergy and the thermal exergy supply from/to the air, nearing the
zero-CO2 emissions responsibility. All of these additional design concepts, systems, and
equipment, when applied collectively, will significantly reduce the total exergy destruction, all
exhibited in the sample calculations. Calculations show that the nearly-avoidable CO2
emissions are low-level of only about 0.05 kg CO2/kW-h of heating or 0.07 CO2/kW-h of
cooling. These values, on an all-season basis, are about one-tenth of the baseline value of 0.63
kg CO2/kW-h [35]. This feature is further enhanced by the CCS system, which may make this
greenhouse a negative-carbon system. This system is also expected to be scaled up to a slow-
city level, as shown in Figure 13. Therefore, it conveys the Agri-voltaic concept to higher and
broader levels for both greener greenhouses and greener cities, being coupled together.

NOMENCLATURE
c Nearly-Avoidable Thermal Exergy Destruction Offset Coefficient,
kg/[kW/kW]
cK Unit emission coefficient of fuel, kg CO2/kW-h
CO2 Direct emission, kg CO2/kW-h
COP Coefficient of Performance
COPEX Exergy-Based Coefficient of Performance
d, e Coefficients of COP Equation (Equation 11)
E Electric power, kW
EDR Environmental Development Ratio
EXE Electrical Power Exergy, kW
EXH Thermal Power Exergy, kW
f Conversion function from PM2.5 to CO2 emission rate
HPF Heat Pump Performance Metric (Equation 18)
In Solar Flux normal to the solar panel Surface, W/m2
PE Pump (Fan) Power Demand, kW
PC Market Price of Solar Panel, €
PEF Primary energy factor (2.5 in Europe)
PM2.5 Particulate matter density in the air with less than 2.5 micrometers (0.025 mm)
diameter, µg/m³
QH Thermal Power, kW
REX Exergy-based renewables mix in the energy supply stock
T Temperature, K
CO2 Nearly-avoidable emission, kg CO2/kW-h
ΣCO2 Total emission, kg CO2/kW-h
T Temperature Difference, temperature drop, K
V Volume flow rate, m3/h
XHP Tret/Tsup
ε Unit exergy, kW/kW or kW-h/kW-h
ψR Rational Exergy Management Efficiency
Subscripts
a Air (Indoor)
B Biogas
BL Boiler
dem Demand
des Destroyed

17
0032-18

E Useful Output, Exit, Electric


EX Exergy
f Fuel, Energy Source
FAN Fan
g Geothermal, ground
H Heat
HE+F Heat exchanger plus fan and motor
in Inlet (temperature) to the indoors
net Net Output
m Mix, mean
o, out Outdoor (Air)
p Panel (solar panel or Radiant Panel)
PV Photovoltaic (Cell)
ref Environment Reference (Temperature)
r, ret Return
sup Supply
X Exergy
Acronyms
AC Alternating Current
ASHP Air-Source Heat Pump (Air-to-air)
ATES Aquifer Thermal Energy Storage
CCS Carbon Capture and Storage
CHP Combined Heat and Power
DC Direct Current
DHW Domestic Hot Water
DE District Energy
5DE Fifth-Generation District Energy System
EAHE Earth-Air Heat Exchanger
ERV Energy Recovery Ventilation
EU European Union
HWP Radiant Wall Panel
LED Light-Emitting Light Lamp (LED Lamp)
nZEXGH Nearly-Zero Exergy Greenhouse
nZCGH Nearly-Zero Carbon Greenhouse
CHP Micro Combined Heat and Power (<50 kW electric power capacity)
100%RHC 100% Renewable Heating and Cooling
OPV Organic Photovoltaic Cell
PCM Phase-changing Material
PV Photovoltaic Solar Panel
PVT Photo-Voltaic-Thermal Solar Panel
REMM Rational Exergy Management Model
TES Thermal Energy Storage
UHI Urban Heat Island Effect

REFERENCES

[1] Rim, M., Sung, Uk-Joo, Tim, K. 2018. Application of Thermal Labyrinth System to Reduce
Heating and Cooling Energy Consumption, Energies 2018, 11, 2762;
doi:10.3390/en11102762.

18
0032-19

[2] Nayak, S., and Tiwari, G. N. 2008. Energy and Exergy Analysis of Photovoltaic/Thermal
Integrated with a Solar Greenhouse, Energy and Buildings, 40 (2008), pp: 2015–2021.
[3] Nayak, S., and Tiwari, G. N. 2010. Energy Metrics of Photovoltaic/Thermal And Earth Air
Heat Exchanger Integrated Greenhouse For Different Climatic Conditions Of India, Applied
Energy, 87 (2010), pp: 2984–2993.
[4] Du, J., Bansal, P., and Huang, B. 2012. Simulation Model of A Greenhouse With A Heat-
Pipe Heating System, Applied Energy, 93 (2012), pp: 268–276.
[5] Ozgener, O., and Ozgener, L. 2011. Determining The Optimal Design of a Closed Loop
Earth To Air Heat Exchanger For Greenhouse Heating By Using Exergoeconomics, Energy
and Buildings, 43 (2011), pp: 960-965.
[6] Kilkis, I. B. 2014. An Exergetic Approach to the Age of Universe, Int. J. Exergy, Vol. 15,
No: 1, pp: 76-89.
[7] Hepbasli, A. 2013. Low Exergy Modelling and Performance Analysis of Greenhouses
Coupled to Closed Earth-To-Air Heat Exchangers (EAHEs), Energy and Buildings, 64 (2013),
pp: 224-230.
[8] Zeiler W., and Boxem, G. 2009. Geothermal Active Building Concept, In book:
Sustainability in Energy and Buildings, Proceedings of the International Conference in
Sustainability in Energy and Buildings (SEB’09), Lee, Shaun H. (Ed.),
ODI: 10.1007/978-3-642-03454-1_32
[9] Schröder D., 2002, Betonkernkühlung mit Zuluft, Besser konditionieren mit weniger
Energieverbrauch, Heizung Lüftung/Klima Haustechnik, Heft 3(2002), seite 47-54.
[10] Kiefer C., 2003, Erfahrungen aus einem Rekord-Sommer, Betonkernkühlung mit Zuluft,
Technik am Bau, 12/2003.
[11] Kilkis, B. 2021. The Relationship Between Climate Emergency, Pandemics, and
Buildings, Manuscript submitted for publication at BRIQ J.
[12] ARUP. 2021. Energy Academy, Groningen, BREEAM Outstanding Label for The
International Research Institute, <Low Energy Building - Energy Academy - Arup>
Last visited on: 24.05.2021
[13] Rotta Loria AF, The Thermal Energy Storage Potential Of Underground Tunnels
Used as Heat Exchangers, Renewable Energy, <https://doi.org/10.1016/j.renene.2021.05.076>.
Last visited on: 24.05.2021
[14] Song, S., Song, J., and Lim J. 2014. Effectiveness of A Thermal Labyrinth Ventilation
System Using Geothermal Energy: A Case Study of An Educational Facility in South Korea,
Energy for Sustainable Development, 23 (2014), pp: 150–164.
[15] Marmion, P., Pradinuk, R., Woods, A., Guity, A., and Rehmanji, I. 2012. Large Dynamic
Thermal Labyrinth: A Step Towards Net Zero Energy Use in Acute Care Hospitals 2012
ACEEE Summer Study on Energy Efficiency in Buildings, 13-215-13-226.
<https://www.aceee.org/files/proceedings/2012/data/papers/0193-000399.pdf> Last visited on: 24.05.2021
[16] Rim, M., Uk-Joo Sung, and Taeyeon Kim. 2018. Application of Thermal Labyrinth
System to Reduce Heating and Cooling Energy Consumption, Energies, 2018, 11, pp: 2762;
DOI:10.3390/en11102762
[17] Misra, R. Jakhar, S., Agrawal, K. K., Sharma, S., Jamuwa, D. K., Soni, M. S., and
Agrawal, D.G. 2018. Field Investigations to Determine The Thermal Performance Of Earth
Air Tunnel Heat Exchanger With Dry And Wet Soil: Energy And Exergetic Analysis, Energy
and Buildings, 171 (2018), pp: 107–115.
[18] EU Science Hub. 2021. Energy Use in The EU Food Sector: State of Play and
Opportunities for Improvement.
[19] World Bank. 2017. SEAR, Energy Access Food and Agriculture,
<http://documents1.worldbank.org/curated/en/417941494928698197/pdf/115062-BRI-P148200-PUBLIC-
FINALSEARSFFoodandAgrigcultureweb.pdf> Last Visited on 22.05.2021.
[20] EU, EUROSTAT. 2020. Agri-environmental Indicator-Energy Use.

19
0032-20

<https://ec.europa.eu/eurostat/statistics-explained/index.php/Agri-environmental_indicator_-_energy_use>
Last Visited on 22.05.2021.
[21] DALSEM. 2021. X-Air Sustainable Greenhouse Solution, <https://www.x-air.nl/>
Last Visited on 22.05.2021.
[22] Kilkis, S. 2014. Energy System Analysis of a Pilot Net-Zero Exergy District, Energy
Conversion and Management, 87 (2014), pp: 1077-1092.
[23] Kilkis, B. 2021. Net-Zero Solar Buildings, What Are They and What They Should Be?
Poster Paper (under Review, SDEWES 2021, Dubrovnik)
[24] Yano, A., Onoe, M., and Nakata, J. 2014. Prototype Semi-Transparent Photovoltaic
Modules for Greenhouse Roof Applications, Biosystems engineering, 122 (2014), pp: 62-73.
[25] Kilkis, B. 2020. Exergy-Optimum Coupling of Heat Recovery Ventilation Units with
Heat Pumps in Sustainable Buildings, J. sustain. dev. energy water environ. syst., 8(4), pp:
815-845. 2020. <DOI: https://doi.org/10.13044/j.sdewes.d7.0316> Last Visited on 22.05.2021.
[26] Weather & Climate. 2021. Climate in Zwolle (Overijssel), Netherlands,
<https://weather-and-climate.com/average-monthly-Rainfall-Temperature-Sunshine,zwolle-overijssel-nl,Netherlands>
Last Visited on 22.05.2021.
[27] Weather & Climate. 2021. Average Monthly Hours of Sunshine In Zwolle (Overijssel)
<https://weather-and-climate.com/average-monthly-hours-Sunshine,zwolle-overijssel-nl,Netherlands>
Last Visited on 22.05.2021.
[28] Godschalk, B., and Bakema, G. 2019. Aquifer Thermal Energy Storage in the Netherlands,
a research programme (2010-2012) Archieving More with Underground Thermal Energy
Storage. Extended English Summary, Book, November 2019 <Meer Met Bodemenergie>
Last Visited on 22.05.2021.
[29] Kilkis, I. B. 2000. Rationalization and Optimization of Heating Systems Coupled to
Ground Source Heat Pumps, ASHRAE Transactions, Vol. 106, Pt. 2, pp: 817-822, 2000,
MN-00-13-1.
[30] REHAU, <https://www.rehau.com/uk-en/ground-heat-exchanger> Last Visited on 22.05.2021.
[31] Ozgener, L., 2011. A Review on The Experimental and Analytical Analysis Of Earth To
Air Heat Exchanger (EAHE) Systems in Turkey, Renewable and Sustainable Energy Reviews,
Elsevier, Vol. 15 (9), pp: 4483-4490.
[32] Kilkis, B. 2021. The Black Sea: A Sea of Energy, Prosperity, and Peace, Belt and Road
Initiative Quarterly, BRIQ Journal, Vol.2, Issue 2, pp: 58-69.
[33] Zorz, J., Richardson, W. D. L., Laventure, A., Bergerson, J., Greogry, W., and Strous, M.
2021. Light Manipulation Using Organic Semiconducting Materials for Enhanced
Photosynthesis, Cell Reports Physical Science, Vol. 2, Issue 4, 100390, April 21, 2021.
<https://www.cell.com/cell-reports-physical-science/fulltext/S2666-3864(21)00080-
1?utm_campaign=STMJ_139381_SC&utm_medium=email&utm_acid=27296758&SIS_ID=&dgcid=STMJ_139381_SC&C
MX_ID=&utm_in=DM149342&utm_source=AC_#%20> DOI: https://doi.org/10.1016/j.xcrp.2021.100390
[34] Wand, D., Liu, H., Li, Y., Lu, X., Chen, H., and Zhi-L, C. 2021.High-Performance and
Eco-Friendly Semitransparent Organic Solar Cells for Greenhouse Applications, Joule, Vol. 5,
Issue 4, pp: 945-957.
<https://www.cell.com/joule/fulltext/S2542-4351(21)00087-
8?utm_campaign=STMJ_139381_SC&utm_medium=email&utm_acid=27296758&SIS_ID=&dgcid=STMJ_139381_SC&C
MX_ID=&utm_in=DM149342&utm_source=AC_>DOI:https://doi.org/10.1016/j.joule.2021.02.010
[35] Kilkis, B. 2021, An Exergy-Based Minimum Carbon Footprint Model for Optimum
Equipment Oversizing and Temperature Peaking in Low-Temperature District Heating
Systems (Under review), Energy.

20

You might also like