Sie sind auf Seite 1von 49

Journal Pre-proof

Energy saving in a convective dryer by using novel real-time exergy-based control


schemes adjusting exhaust air recirculation

Saman Zohrabi, Mortaza Aghbashlo, Seyed Sadegh Seiiedlou, Holger Scaar, Jochen
Mellmann

PII: S0959-6526(20)30441-8
DOI: https://doi.org/10.1016/j.jclepro.2020.120394
Reference: JCLP 120394

To appear in: Journal of Cleaner Production

Received Date: 19 October 2019


Revised Date: 31 January 2020
Accepted Date: 1 February 2020

Please cite this article as: Zohrabi S, Aghbashlo M, Seiiedlou SS, Scaar H, Mellmann J, Energy saving
in a convective dryer by using novel real-time exergy-based control schemes adjusting exhaust air
recirculation, Journal of Cleaner Production (2020), doi: https://doi.org/10.1016/j.jclepro.2020.120394.

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition
of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of
record. This version will undergo additional copyediting, typesetting and review before it is published
in its final form, but we are providing this version to give early visibility of the article. 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.

© 2020 Published by Elsevier Ltd.


Saman Zohrabi: Investigation; Software; Data Curation

Mortaza Aghbashlo: Conceptualization; Writing - Review & Editing

Seyed Sadegh Seiiedlou: Supervision; Conceptualization

Holger Scaar: Resources; Methodology

Jochen Mellmann: Resources; Writing - Review & Editing

1
1 Energy saving in a convective dryer by using novel real-time exergy-

2 based control schemes adjusting exhaust air recirculation

4 Authors:

5 Saman Zohrabia,b, Mortaza Aghbashloc,*, Seyed Sadegh Seiiedloua,*,

6 Holger Scaarb, Jochen Mellmannb

7 Affiliation:
a
8 Department of Biosystems Engineering, University of Tabriz, Tabriz, Iran
b
9 Department of Postharvest Technology, Leibniz Institute for Agricultural Engineering
10 and Bioeconomy, Potsdam, Germany.
c
11 Department of Mechanical Engineering of Agricultural Machinery, Faculty of
12 Agricultural Engineering and Technology, College of Agriculture and Natural Resources,
13 University of Tehran, Karaj, Iran.
14

15

16

17 Corresponding author:

18 Mortaza Aghbashlo, maghbashlo@ut.ac.ir

19 Seyed Sadegh Seiiedlou, seiidloo@tabrizu.ac.ir

20

21

22

23

24

1
25 Abstract

26 This study was aimed at developing and comparing three different novel real-time exergy-

27 based control schemes for adjusting exhaust air recirculation in a convective dryer for the

28 first time. These control schemes were based on output–input warm exergy ratio of drying

29 cabinet (Scheme I), output–input warm exergy ratio of drying cabinet considering a fixed

30 upper limit for the output wet exergy rate (Scheme II), and output–input warm exergy ratio of

31 drying cabinet considering a varying upper limit for the output wet exergy rate (Scheme III).

32 A continuous time-based control approach (Scheme IV) was also developed for comparison

33 purposes. The exergetic effectiveness improvement and energy saving of the process

34 achieved using the developed control schemes were assessed with respect to the base

35 operation mode (without exhaust air recirculation) in order to make decisions on the most

36 efficient approaches. As a case study, poplar wood chips were dried at two temperatures (55

37 and 70 ºC) and two volumetric flow rates (360 and 450 m3/h) of drying air. The exergetic

38 effectiveness of the drying process varied from 1.4 to 3.1%. The experimental results showed

39 that the exergetic effectiveness of the drying process could be improved by employing all the

40 developed control schemes. Among the developed control schemes, the control scheme III

41 showed the highest exergetic effectiveness improvement in the range of 46.5‒75.9%. This

42 control scheme could also reduce the overall energy consumption of drying system in the

43 range of 30.7‒34.5% compared with the base operation mode. The selected control scheme

44 could be implemented on various drying systems in order to improve their efficiency,

45 productivity, and sustainability by recuperating the waste exergy from exhaust air.

46
47 Keywords: Convective drying; Exergetic effectiveness; Exhaust air recirculation; Poplar

48 wood chips; Real-time control; Output-input warm exergy ratio

49

2
50 Research highlights:

51 - Exhaust air recirculation of a convective air dryer was exergetically controlled.

52 - The developed control schemes improved the exergetic effectiveness of the process.

53 - The optimal control method improved the exergetic effectiveness of the process up to 75%.

54 - The selected control technique reduced the overall energy needs of the process up to 34%.

55

Notations Subscripts
 Cross-sectional area ( ) 0 Reference state
 Specific heat capacity (/
) Air
 Energy rate ( or /)  destruction
 Exergy rate ( or /)  Equilibrium
ℎ Latent heat of vaporization at saturation state (/
)  Evaporation
 Mass flow rate (
/)  loss
 Moisture content (
 /
  ) ! Initial
" Moisture ratio (-) # Product
# Probability value (p-value)  Drying time
$ Pressure ($ ) ! Total
% Volumetric air flow rate (& ⁄ℎ) 
% Heat flow rate ()
Vapor
 Saturated vapor
" Gas constant (⁄
)  Water
* Temperature ( )  Wet
+ Velocity (/) m
 Work rate ()
Warm

Greek letters
Θ Output–input warm exergy ratio of drying cabinet (-)
- Density (
/& )
. Relative air humidity (%"0)
1 Exergetic effectiveness improvement of drying system (-)
2 Exergetic effectiveness of drying process (%)
3 Absolute air humidity (
 ⁄
4 5)
56

57

58

59

60

61

62

63

64

65

66
3
67 1. Introduction

68 Convective air drying is a highly demanding unit operation in manufacturing industry

69 due to its low capital and operating costs. This versatile drying technique is used in

70 various industries for removing moisture from wet materials. The convective drying process

71 consumes a huge amount of thermal energy, frequently produced through burning fossil fuel,

72 for vaporizing water. According to Perry’s Chemical Engineers’ Handbook, the specific

73 thermal energy consumption of this drying method has been reported to be in the range of

74 5,000–7,350 kJ per kilogram of evaporated water (Perry et al., 2008). In order to minimize

75 the energy requirements of convective drying systems, several different approaches like

76 exhaust air recirculation, waste heat exchanging, and heat pump recovering have been

77 proposed and implemented on various drying systems (Ghasemkhani et al., 2016).

78 Exhaust air recirculation is the simplest and most widely used approach of recovering

79 waste heat in convective dryers due to its ease of implementation and low cost. However, full

80 recirculation of exhaust air in drying systems might demand additional energy in order to

81 dehumidify the moist outflow air (Mandegari and Pahlavanzadeh, 2013). Hence, partial

82 recirculation of exhaust air without the need for its dehumidification has been appeared to be

83 a technically simple and economically viable strategy for recuperating drying potential of the

84 air leaving drying system. In the light of this, various research works have been devoted so

85 far to determining the potential energy saving and thermodynamic efficiency improvement of

86 drying systems achieved through recirculating the exhaust air. For example, Tippayawong et

87 al. (2009) introduced a recirculating air dryer for improving energy use in traditional longan

88 fruit drying. Golman and Julklang (2014) applied model simulations to study the effects of

89 process variables on the energy recovery in spray drying process. According to the published

90 literature, partial recirculation of exhaust air can markedly lower the energy requirements of

91 drying systems.
4
92 Despite the fact that the use of partial recirculation of exhaust air is not new in drying

93 technology, it is believed that advanced engineering solutions should still be introduced and

94 implemented for optimizing and controlling the recirculation ratio. The optimized

95 recirculation ratio not only can recover the major portion of thermal energy available in the

96 exhaust air but also can maintain the quality of dried product. Usually, decisions on the

97 recirculation ratio are made on the basis of temperature and humidity (enthalpy) of exhaust

98 air. Furthermore, the performance improvement potential of drying systems is often

99 expressed in terms of thermal efficiency. However, temperature and humidity (enthalpy) of

100 the outflow air cannot well reflect its real drying capacity. In addition, energy-based

101 performance indices might propel decision-makers towards wrong decisions on the

102 recirculation ratio (Mkwananzi et al., 2019). These are attributed to the fact that neither

103 temperature and humidity values nor quantity energy can value the true capacity of a hot air

104 stream for doing evaporation or performing useful work. This issue can be effectively tackled

105 by the use of exergy concept in which the true capacity of an energy stream for doing useful

106 work can be reliably valued (Aghbashlo and Rosen, 2018). This engineering measure

107 depends on not only the thermodynamic properties of the system but also the specifications of

108 the reference system, often considered the ambient environment (Aghbashlo et al., 2017).

109 During the past decades, various tools like energy, emergy, life cycle assessment, and

110 exergy approaches have been developed for assessing the environmental and sustainability

111 aspects of energy systems (Aghbashlo et al., 2018c). Among these techniques, exergy

112 analysis is an universally accepted tool for analyzing energy systems from environmental and

113 sustainability viewpoints since it is developed on the basis of a well-defined methodology

114 (Gasparatos et al., 2008). More specifically, the exergy concept can precisely measure the

115 environmental and sustainability performance of energy systems due to the fact that it is

116 rooted in engineering thermodynamics (Aghbashlo et al., 2018a). This issue has also been

5
117 documented by Dincer and Rosen (2005) as portrayed in Figure 1. Obviously, increasing

118 exergy efficiency of a given energy system can directly enhance its sustainability level while

119 mitigating the environmental impact associated with the system and vice versa (Aghbashlo et

120 al., 2018b).

121

122
123 Figure 1. Association between environmental impact and sustainability level of a given
124 energy system with its exergy efficiency (Dincer and Rosen, 2005). With permission from
125 Elsevier. Copyright© 2020.
126

127 These are why the exergy concept has been widely adopted for analyzing, optimizing, and

128 retrofitting various drying processes and systems. For example, exergy analysis has been used

129 for the evaluation of potato slices drying (Aghbashlo et al., 2008), carrot slices drying

130 (Aghbashlo et al., 2009), hot air drying of olive leaves (Erbay and Icier, 2009), carrot cubes

131 fluidized bed drying (Nazghelichi et al., 2010), spray drying of white cheese slurry (Erbay

132 and Koca, 2012), industrial pasta drying (Colak et al., 2013), rough rice fluidized bed drying

6
133 (Khanali et al., 2013), potato crisp frying (Genc and Hepbasli, 2015), spray drying of whole

134 milk (Yildirim and Genc, 2017), fluidized bed drying of paddy (Sarker et al., 2015), turmeric

135 (Curcuma longa) solar drying (Karthikeyan and Murugavelh, 2018), pistachio drying and

136 roasting (Sheikhshoaei et al., 2019), and solar-infrared drying of mint and apple slices (Şevik

137 et al., 2019).

138 Even though a number of studies have tried to analyze and optimize various

139 drying systems using the exergy concept as listed-above, those investigations were only

140 aimed at identifying the breakthrough points for further exergetic improvements. Literature

141 survey showed that the use of exergy concept for real-time control of exhaust air recirculation

142 in drying technology has not been introduced and experimented yet. In fact, the majority of

143 the work conducted thus far on controlling and adjusting the recirculation ratio has been

144 based on temperature, humidity, and enthalpy of the air leaving drying chamber.

145 Accordingly, the goal of this study was to introduce three novel real-time exergy-based

146 control schemes for exhaust air recirculation and to examine their capabilities for boosting

147 the performance of a convective dryer used for poplar wood chips drying for the first time. It

148 is noteworthy that this study is a continuation of the authors’ previous study (Zohrabi et al.,

149 2019) in which the exergetic performance of the dryer was markedly improved via exhaust

150 air recirculation at a constant fraction. In addition, the exergetic content of the outflow air

151 was split into two portions, i.e., warm exergy and wet exergy. Overall, it was concluded that

152 decisions on the recirculation ratio should be made with respect to the warm exergetic content

153 of the outflow air rather than its total exergetic value. Accordingly, this study was conducted

154 to optimize and control the ratio of exhaust air recirculation using three real-time control

155 schemes based on output–input warm exergy ratio of drying cabinet along with a fixed and

156 varying upper limit for the output wet exergy rate of drying cabinet. The capabilities of the

157 developed exergy-based control schemes for boosting the exergetic performance and

7
158 decreasing the energy consumption rate of the dryer were also compared with those of base

159 operation mode and continuous time-based control approach.

160

161 2. Materials and methods

162 2.1. Experimental set-up

163 A pilot-scale convective dryer equipped with a real-time air recirculation adjustment

164 system was used in this study. The drying system was developed at the Leibniz Institute of

165 Agricultural Engineering and Bio-economy (ATB) in Potsdam, Germany. It is schematically

166 depicted in Figure 2. Detailed information regarding the instrumentation and operation of the

167 dryer can be found in Zohrabi et al. (2019).

168
169 Figure 2. Schematic illustration of the developed convective dryer.
170

171 Figure 3 shows a schematic illustration of the process flow diagram of drying. The drying

172 process proceeded in the following order: drying cycle was started by intaking dry ambient

8
173 air using a centrifugal fan (points 1‒4). The intake air was heated using an electrical air heater

174 (points 4–5). Both centrifugal fan and electrical air heater were controlled by PI controllers.

175 The heated air was blown through the drying cabinet where the drying process took place

176 (point 5‒6). A portion of the moist drying medium exhausting from the drying cabinet was

177 recirculated based on the control strategy chosen using three different flaps as shown in

178 Figure 2. The rest of the air was vented to the atmosphere via an exhaust duct as indicated in

179 Figure 3 (point 6‒8).

180 Various measuring sensors were placed at different points of the dryer in order to monitor

181 and record temperature, velocity, relative humidity, and pressure drop of drying air in a

182 continuous manner (Figure 3). Furthermore, the drying cabinet mass as well as the electrical

183 energy consumption rate were measured in the course of drying process. Data sampling and

184 process control were conducted using a 4-Slot, USB CompactDAQ Chassis (NI cDAQ-9174)

185 through LabVIEW® (R15.0) software.

186

187
9
188 Figure 3. Schematic illustration of the drying process flow diagram. 1) Intake air; 2)
189 Intake-recirculated air mixture; 3) Filtered clean air; 4) Blown pressurized air; 5) Heated air;
190 6) Outflow moist air; 7) Vent air; 8) Recirculated air; 9) Wet product; and 10) Dried product.
191

192 2.2. Drying experiments

193 Two temperatures (55 and 70 °C) and two volumetric flow rates (360 and 450 m3/h) of

194 drying air were taken into consideration herein for drying experiments. Wood chips obtained

195 from poplar plantations were used as test material. At the inception of each

196 drying experiment, the dryer was run for about 30 min in order to achieve the desired air

197 velocity and temperature. Once the dryer stabilized, 10 kg of wood chips was transferred to

198 the drying cabinet. All experimental trials were replicated twice. The drying process was

199 terminated as soon as the moisture ratio of the sample decreased to about 0.1. The moisture

200 ratio of the sample in the course of drying was determined as


78 97:;
" =
7< 97:;
(1)
201 where = stands for the moisture content (kg water/kg dry matter) at a given time. > and

202 ?@ are the initial and equilibrium moisture contents (kg water/kg dry matter), respectively.

203 Notably, the equilibrium moisture content (?@ ) for the poplar wood chips with a high initial

204 moisture content could be ignored (Tzempelikos et al., 2014).

205

206 2.3. Developing control schemes

207 Zohrabi et al. (2019) deduced that decision-making on the recirculation ratio with respect

208 to the total exergy rate of the air stream exhausting from drying cabinet might lead to

209 misconclusions since increasing the humidity ratio could markedly increase this parameter.

210 Using the concept of warm and wet exergy elaborated by Shukuya (2012), Zohrabi et al.

211 (2019) concluded that warm exergetic content of the air leaving drying cabinet should be

212 taken into account for the right decision making on the recirculation ratio. Notably, Shukuya

10
213 (2012) divided the total exergy content of a moist air stream into two portions, i.e., warm and

214 wet exergies. In a typical drying process, two forces act against each other, i.e., warm exergy

215 (as the driving force for mass transfer) and wet exergy (as the opposing force for mass

216 transfer). In better words, the higher the warm exergy, the higher is the drying rate and vice

217 versa. According to the above arguments, output–input warm exergy ratio (Θ) of drying

218 cabinet as defined in Eq. 2 could be used for decision-making on the recirculation ratio of the

219 exhaust air.

BCDE,G
A=
BCDE,H
(2)

220 where  is the exergy rate and the subscript  shows the warm exergy.

221 The warm exergy content of an air stream could be calculated as follows (Zohrabi et al.,

222 2019):

I,JK = I LM( )I + 3I ( )O P(*I − *R )


*I $I
− *R SM( )I + 3I ( )O P L T − ("I + 3I "O ) L TUT
(3)
*R $R
223 where  denotes the mass flow rate,  represents the specific heat capacity, 3 stands

224 for the humidity ratio, * indicates the temperature, " is the gas constant, and $ shows the

225 pressure. The subscripts , , and 0 refer to the air, vapor, and reference state, respectively .

226 The wet exergy content of an air stream could be determined using the following equation

227 (Zohrabi et al., 2019):

1 + 1.60783R 3I
I,J= = I L*R S("I + 3I "O ) L T + 1.60783R "I  L TUT
1 + 1.60783I 3R
(4)

228 where the subscripts  refers to the wet exergy.

229 Accordingly, the total exergy rate of an air stream could be determined using the equation

230 given below (Mojarab Soufiyan et al., 2016):

11
I = I,JK + I,J=
= I LM( )I + 3I ( )O P(*I − *R )
*I $I
− *R SM( )I + 3I ( )O P L T − ("I + 3I "O ) L TU
*R $R
(5)
1 + 1.60783R
+ *R S("I + 3I "O ) L T
1 + 1.60783I
3I
+ 1.60783R "I  L TUT
3R
231 The equations and constants used in the calculation of warm and wet exergies of air

232 streams are tabulated in Table 1.

233
234 Table 1. Equations and constants used in the calculation of warm and wet exergies of air
235 streams.
Parameter Equation/Constant* Ref.
-I %I
I = = -I +I \]^=
3600
Mass flow rate of air -

Specific heat capacity of dry _ `I = 1006.4 + 1.15 × 109 *I + 5.0 × 109d *I − 8.0 × 109e *I& (Martin et al.,
air ‒120≤T(ºC) ≤140 2002)

_ `O = 1887.2945 + 0.81825992*I + 7.5620368 × 109& *I&


− 2.7868655 × 109m *I& + 6.1460039 × 109e *Id
Specific heat capacity (Martin et al.,
‒0.01≤T(ºC) ≤150
of water vapor 2002)

.$On,Im
3I, = 3I,& = 3I,d = 3I,m = 0.622
$R − .$On,Im
Absolute air humidity -

6887 *I,m
$On,Im = 0.1# o27.014 − − 5.31 L Tp
*I,m 273.16
Saturated vapor pressure of (Ranjbaran et al.,


ambient air 2014)

3I,q I + (J )?O


3I,q = 3I,e = 3I,r =
I
Absolute air humidity -
Mass flow rate of the (J )?O = J,s − J,tR -
vaporized water
0.49 m2
Drying cabinet cross-
-
sectional area (A)
Reference temperature (*R ) 20 ºC** -
Reference pressure ($R ) 101.3 kPa -
Air density (-) 1.225 kg/m3 -
Gas constant for air ("| ) 0.287 kJ/kg K -
0.4615 kJ/kg K
Gas constant for water vapor
-
(" )
0.006480 kg water/kg dry air
Absolute air humidity of
-
reference state (3R )
236 *where -, %, +, , and . respectively denote for the density, volume flow rate, air velocity, surface area, and relative
237 humidity. The subscripts ‡ˆ, , , and  respectively refer to drying duct, saturated vapor, water, and evaporation.
238 ** Despite the fact that the reference temperature is often assumed to be 25 °C in the exergetic studies (Yang et al.,
239 2019a, 2019b), the average ambient temperature (i.e., 20 °C) during the period of experiments was used as a reference
240 temperature herein in order to make the obtained exergetic results realistic as much as possible.
241

242 The developed control schemes for adjusting the recirculation ratio of the exhaust air were
12
243 as follows:

244 • Scheme I– Output–input warm exergy ratio of drying cabinet; the exhaust air from

245 drying cabinet was recirculated in proportion to the magnitude of output–input warm

246 exergy ratio.

247 • Scheme II– Output–input warm exergy ratio of drying cabinet considering a fixed

248 upper limit for the output wet exergy rate; the air exhausting from drying cabinet was

249 recirculated in proportion to the magnitude of output–input warm exergy ratio when

250 the output wet exergy rate was below 0.1 kW. This limitation was chosen based on

251 the experiments previously conducted at a fixed exhaust air recirculation (Zohrabi et

252 al., 2019).

253 • Scheme III– Output–input warm exergy ratio of drying cabinet considering a varying

254 upper limit for the output wet exergy rate; the air exhausting from drying cabinet was

255 recirculated in proportion to the magnitude of warm exergy ratio when the output wet

256 exergy rate was lower than the average wet exergy rate obtained in the previous 30

257 minutes intervals for the base operation mode (without air recirculation).

258 • Scheme IV– Continuous time-based control approach; the exhaust air from

259 drying cabinet was recirculated in proportion to the time of drying process

260 obtained previously for the base operation mode .

261 It should be noted that the recirculation ratio was continuously adjusted during the process

262 in a real-time manner through calculating the output–input warm exergy ratio and,

263 subsequent, changing the flaps position. The recirculation ratio vs. output–input warm exergy

264 ratio for the first control scheme is shown in Figure 4. The recirculation ratio vs. output–input

265 warm exergy ratio for the second control scheme was similar to the first control scheme

266 unless the output wet exergy rate went beyond 0.1 kW. When the output wet exergy rate

267 exceeded 0.1 kW, the recirculation flaps were stepped back in one-minute intervals
13
268 automatically until the output wet exergy rate fell down again below the upper limit of 0.1

269 kW. The recirculation ratio vs. output–input warm exergy ratio for the third control scheme

270 was similar to the first control scheme unless the output wet exergy rate exceeded the average

271 output wet exergy rate obtained in the previous 30 minutes intervals for the base

272 operation mode. When the output wet exergy rate went beyond the average output wet exergy

273 rate, the recirculation flaps were stepped back in one-minute intervals automatically until the

274 output wet exergy rate dropped again below the average output wet exergy rate. In the fourth

275 control scheme, the exhaust air from drying cabinet was proportionally recirculated in a

276 continuous manner with respect to the time of drying process obtained previously for the base

277 operation mode. The real-time control flow chart used herein for adjusting the recirculation

278 ratio is shown in Figure 5.

279
1.0
0.9
Recirculation ratio (-)

0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0

280 Output-input warm exergy ratio (-)


281 Figure 4. The recirculation ratio vs. output–input warm exergy ratio for the first control
282 scheme.

14
283
284 Figure 5. Real-time control flow chart used for adjusting the recirculation ratio.
285
286
287 Improvement in the exergetic effectiveness of drying process (2) obtained using the

288 developed control schemes was determined with respect to the base operation mode in order

289 to measure their usefulness.

15
Š D‹Œ ŽŒ‘’ − Š D‹Œ“Œ ŽŒ‘’
‰=
Š D‹Œ“Œ ŽŒ‘’
(6)

290 The above-mentioned index was designated as “exergetic effectiveness improvement” (1).

291 In the equation above, the subscripts 5ℎ ˆ!! and 5ℎ!‡ ˆ!! refer to the process

292 with and without control, respectively.

293 The exergetic effectiveness of drying process is the ratio of output, the exergy sought, to

294 input, the exergy that costs. The following equation could be used for computing the

295 exergetic effectiveness of drying process for each experimental trial.

BC”• + _BC–,—˜ − BC–,™ `


Š=
BCš,— − BCš,› + œŒŒ
(7)

296 where  stands for the work (electricity) rate as well as the subscripts # and ! refer to

297 the product and total, respectively. It is worth noting that decreasing exergy loss from drying

298 system could discount the electricity consumption rate which in turn could lead to an increase

299 in the exergetic effectiveness of drying process and vice versa.

300 The equations and constants used in the calculation of the exergetic effectiveness of drying

301 process are tabulated in Table 2.

302
303 Table 2. Equations and constants used in the calculation of exergetic effectiveness of drying
304 process.
Parameter Equation/constant* Ref.
*R
?O = 1 − ž %?O
The rate of exergy utilized for water (Sheikhshoaei et al.,
evaporation from the product * 2019)
%?O = (J )?O ℎ
(Sheikhshoaei et al.,
Heat flow rate as a result of evaporation
2019)
ℎ = 2.503 × 10q − 2.386
× 10& (* − 273.16) (Brooker et al.,
Latent heat of water evaporation at
273.16≤T(K) ≤338.72 1992)
ℎ = (7.33 × 10t − 1.60 × 10e *  )R.m
saturation state
338.72≤T(K) ≤533.16
 =  _ ` Ÿ_*  − *R `
* 
(Yildirim and Genc,
Exergy rate of the poplar wood chips
− *R   ž¡
2015)
*R
Specific heat capacity of the poplar _ ` = 0.1031 + 0.003867*  (Coskun et al., 2009)
wood chips
305 *where % and ℎ respectively denote the heat flow rate and latent heat of vaporization at saturation state.

16
306 It should be noted that changes in the wood chips composition were neglected in the

307 exergetic calculations. The error imposed by this simplification was negligible since the

308 exergy rate of the poplar wood chips was markedly lower than the other exergetic terms

309 considered in the computations.

310 The following exergy balance equation was used to quantify the rate of exergy destruction

311 of the drying system operating under different control schemes:

BCš,— + BC–,™ + œŒŒ = BCš,› + BC–,—˜ + BC”• + BC’ + BC¢”£ (8)

312 where the subscripts  and  respectively refer to loss and destruction.

313 The rate of exergy loss from the drying system was determined as follows (Aghbashlo et

314 al., 2019):

¤˜
BC’ = L— − T ¥’
¤’
(9)

315 The rate of energy loss from the drying system was determined as follows:

 š,— + B
B  –,™ + œŒŒ = B  –,—˜ + ¥”• + ¥’
 š,› + B (10)

316 where  stands for the energy rate.

317 The energy rates of drying air and product were obtained using the following equation,

318 respectively:

I = I ¦M( )I + 3I ( )O P(*I − *R )§ (11)

 =  _# `# M_*  − *R `P (12)

319

320 2.4. Uncertainty and statistical analyses

321 The methodology elaborated by Holman (2001) was applied for determining the

322 uncertainties associated with experimental data and calculated parameters. The uncertainties

323 related to the measured data and exergetic parameters were found to be below 5% which

17
324 were acceptable for an experimental study (Colak and Hepbasli, 2007).

325 The experimental results were statistically evaluated using a 2×2×5 split factorial design

326 (two temperatures and two volumetric flow rates of drying air as well as five control

327 schemes) with two repetitions for each experiment. A one-way ANOVA analysis was carried

328 out to find the effects (p-value < 0.05) of the experimental variables on the exergetic

329 performance parameters of the dryer. Notably, the p-value or probability value lower than

330 0.05 reveals the fact that there is a strong evidence against the null hypothesis and, therefore,

331 the null hypothesis should be rejected. However, the p-value higher than 0.05 shows a weak

332 evidence against the null hypothesis and, therefore, the null hypothesis is true. Tukey’s

333 multiple comparison tests were conducted at 5% level of significance. SPSS (IBM SPSS

334 Statistics 24) software was used for carrying out all the analyses.

335

336 3. Results and Discussion

337 The variations in the moisture ratio of the poplar wood chips as a function of drying time

338 at T=70 ºC and Q=450 m3/h under various control schemes are depicted in Figure 6. Figure 7

339 also illustrates the variations in the moisture ratio of the poplar wood chips as a function of

340 drying time at different experimental conditions under the control scheme III. Obviously, the

341 product moisture content significantly declined with increasing drying time (p < 0.05).

342 Generally, the drying process of wood chips occurred in the falling rate period, manifesting

343 the importance of internal molecular diffusion during drying. According to Figure 6, there

344 was not a significant change (p > 0.05) in drying time of poplar wood chips under various

345 control schemes. In better words, the drying behavior of poplar wood chips under all the

346 developed control schemes was similar to each other. Augmenting temperature and

347 volumetric flow rate of drying air could elevate both heat and mass transfer coefficients

348 which in turn could shorten the time required to reduce the moisture ratio of wood chips to
18
349 about 0.1 (Figure 7).

350

351
1.0
Base operation mode
0.9 Scheme I
0.8 Scheme II
Moisture ratio (-)

0.7 Scheme III


Scheme IV
0.6
0.5
0.4
0.3
0.2
0.1
0.0
0 20 40 60 80 100 120
352 Drying time (min)
353 Figure 6. Variations in moisture ratio of the poplar wood chips as a function of drying time at
354 T=70 ºC and Q=450 m3/h under various control schemes.
355

356
1.0
T=55℃, Q=360 m³/h
0.9
T=55℃, Q=450 m³/h
0.8
T=70℃, Q=360 m³/h
Moisture ratio (-)

0.7
T=70℃, Q=450 m³/h
0.6
0.5
0.4
0.3
0.2
0.1
0.0
0 30 60 90 120 150 180 210
Drying time (min)
357
358 Figure 7. Variations in moisture ratio of the poplar wood chips as a function of drying time at
359 different experimental conditions under the control scheme III.
360

361 The variations in the exhaust air recirculation ratio vs. drying time for the developed

362 control schemes at T=70 ºC and Q=450 m3/h are represented in Figure 8. For the base

363 operation mode, the exhaust air recirculation flap was forced to stay at zero position (closed).

364 For the control schemes I, II, and III, the position of recirculation flap was adjusted according

19
365 to the process conditions. In addition, the flap was continuously opened from zero to 100% in

366 proportion to the time of drying process for the control scheme IV. Clearly, the control

367 scheme III showed more fluctuations in the position of recirculation flap over time mainly

368 caused by controlling the output wet exergy of drying cabinet.

369
100
Base operation mode
90 Scheme I
Exhaust air recirculation ratio (-)

80 Scheme II
70 Scheme III
Scheme IV
60
50
40
30
20
10
0
0 20 40 60 80 100 120
370 Drying time (min)
371 Figure 8. Exhaust air recirculation ratio vs. drying time for the developed control schemes at
372 T=70 ºC and Q=450 m3/h.
373

374 Table 3 summarizes the average thermodynamic parameters of streams of the

375 drying system with respect to their state numbers given in Figure 3. Figure 9 indicates the

376 variations in the output warm, wet, and total exergy rates of drying cabinet at T=70 ºC and

377 Q=450 m3/h under different control schemes. Clearly, the output warm exergy rate of drying

378 cabinet increased linearly (p < 0.05) in a significant manner as drying process proceeded

379 (Figure 9A). This could be related to the fact that the majority of the warm exergy supplied to

380 drying cabinet was not appropriately used for moisture evaporation since the moisture content

381 of the product diminished towards the end of the process. The developed control schemes did

382 not significantly affect (p > 0.05) the output warm exergy rate of drying cabinet. This could

383 be explained by the fact that the input warm exergy rate of drying cabinet was approximately

384 the same for different control schemes at a given drying condition. On the other hand, the

20
385 evaporated water could not accumulate within the drying system under different control

386 schemes because of continuous venting of a portion of drying air. Accordingly, the output

387 warm exergy rate of drying cabinet came close to its input value at a given drying condition.

388 The output wet exergy rate of drying cabinet continuously decreased over time for the

389 base operation mode (Figure 9B). This could be attributed to the fact that the moisture

390 content of the product was decreased by proceeding drying process while the moisture taken

391 from the product was vented to the atmosphere. However, the output wet exergy rate of

392 drying cabinet was increased initially with progressing drying time and was then declined

393 towards the end of the drying process for all the developed control schemes. Unlike the base

394 operation mode, the moisture evaporated from the product was accumulated within drying

395 medium in proportion to the magnitude of exhaust air recirculation ratio. Accordingly, the

396 output wet exergy rate of drying cabinet was significantly higher (p < 0.05) for all the

397 developed control schemes in comparison with the base operation mode. The output wet

398 exergy rate of drying cabinet under the control scheme III was close to the base operation

399 mode. In fact, there was no significant moisture accumulation within drying medium under

400 the control scheme III compared with the base operation mode. This occurred due to the

401 stringent control of output wet exergy rate of drying cabinet under this control method as

402 explained above. In addition, the capability of drying medium for removing water from wood

403 chips under the control scheme III was comparable with the base operation mode. It should

404 be noted that the warm exergy content of a hot air stream is the sole driving force for water

405 removal from the product being dried as explained by Zohrabi et al. (2019).

406 The total output exergy rate of drying cabinet was increased linearly (p < 0.05) in a

407 significant manner over drying time for all the developed control schemes (Figure 9C). It is

408 obvious from Figure 9 that the contribution of warm exergy content to the total output exergy

409 of drying cabinet was dominant compared with that of wet exergy content. In contrast, at a

21
410 constant exhaust air recirculation ratio, the output wet exergy rate of drying cabinet

411 accounted for the major portion of its total output exergy rate (Zohrabi et al., 2019). In fact,

412 the water vapor accumulated within the drying medium at a constant exhaust air recirculation

413 ratio increased its absolute air humidity and, thereby, the output wet exergy rate. However,

414 during the real-time control of exhaust air recirculation ratio, particularly in the control

415 scheme III, a portion of the accumulated water vapor within the drying medium was

416 discharged to the ambient. This in turn decreased the output wet exergy rate of drying

417 cabinet. Generally, the total output exergy rate of drying cabinet under all the developed

418 control schemes was higher than that of base operation mode because of moisture

419 accumulation within drying medium.

420

421

422

423

424

425

426

427

428

22
429 Table 3. Average thermodynamic parameters of streams of the drying system.
Base operation mode Scheme I Scheme II Scheme III Scheme IV
Q=360 m³/h Q=450 m³/h Q=360 m³/h Q=450 m³/h Q=360 m³/h Q=450 m³/h Q=360 m³/h Q=450 m³/h Q=360 m³/h Q=450 m³/h
Parameter Unit
T=55 T=70 T=55 T=70 T=55 T=70 T=55 T=70 T=55 T=70 T=55 T=70 T=55 T=70 T=55 T=70 T=55 T=70 T=55 T=70
°C °C °C °C °C °C °C °C °C °C °C °C °C °C °C °C °C °C °C °C
*I,t °C 21.6 21.9 20.8 21.8 19.1 20.0 19.4 19.7 21.4 20.5 21.5 21.4 20.2 21.2 20.5 20.1 21.9 20.9 21.1 21.9
*I, °C 20.8 23.4 23.3 23.1 30.6 31.8 32.3 36.3 34.3 35.0 37.5 36.2 30.8 33.2 31.0 36.0 34.9 36.6 35.5 37.1
*I,m °C 55.0 70.0 55.0 69.6 55.1 70.1 55.1 67.8 55.0 70.1 55.1 67.5 55.0 70.1 55.2 67.9 55.0 70.2 55.1 69.0
*I,q °C 42.8 50.0 44.0 52.1 41.5 50.7 44.0 51.4 44.0 50.6 46.7 50.5 39.8 50.9 43.6 52.4 44.3 53.5 46.3 52.6
*I,e °C 39.5 46.3 42.1 49.9 40.8 46.2 41.7 48.7 41.8 47.6 45.1 49.7 39.1 46.6 39.8 49.0 42.9 49.3 44.6 49.9
*,s °C 20.0 20.0 20.0 20.0 20.0 20.0 20.0 20.0 20.0 20.0 20.0 20.0 20.0 20.0 20.0 20.0 20.0 20.0 20.0 20.0
*,tR °C 43.0 47.3 44.7 55.7 44.7 46.1 45.1 48.2 38.9 48.4 45.0 51.2 46.2 50.1 48.1 53.5 43.4 46.8 48.6 49.4
I,t kg/h 433.3 433.3 541.7 541.7 224.3 223.6 238.6 256.7 186.9 237.7 182.6 296.4 201.2 233.0 270.8 254.6 185.9 205.5 231.1 252.5
Im kg/h 434.9 436.2 545.3 545.2 436.9 438.0 547.3 548.6 438.9 438.0 547.7 547.6 437.7 437.5 547.2 546.7 437.9 438.6 546.7 548.8
I,q kg/h 436.3 438.1 547.2 547.4 438.2 440.1 549.0 550.9 440.1 440.0 548.7 549.7 439.5 439.6 549.3 548.8 438.9 440.4 547.9 550.9
I,e kg/h 437.6 439.4 549.0 549.1 227.5 227.7 242.6 261.7 190.3 242.1 185.7 301.6 204.6 237.1 275.2 258.9 188.9 209.5 234.6 257.4
$I,t kPa 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3
$I,m kPa 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.4
$I,q kPa 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.4 101.3 101.4 101.4 101.4 101.4 101.4 101.4 101.4
$I,e kPa
I,t
101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3 101.3
kJ/s
I,
0.203 0.240 0.124 0.288 0.009 0.007 0.032 0.004 0.073 0.027 0.042 0.089 0.014 0.074 0.043 0.012 0.099 0.048 0.045 0.131
kJ/s
I,m
0.097 0.417 0.507 0.479 1.384 1.504 1.924 2.561 1.794 1.882 2.738 2.542 1.353 1.643 1.713 2.490 1.865 2.083 2.406 2.687
kJ/s
I,q
3.709 5.201 5.016 7.020 3.950 5.512 5.147 6.972 4.074 5.577 5.200 6.906 3.998 5.516 5.150 6.838 4.072 5.660 5.192 7.169
kJ/s
I,e
2.823 3.729 3.720 4.993 2.676 3.861 3.754 4.954 3.024 3.852 4.181 4.796 2.484 3.883 3.708 5.072 3.039 4.216 4.094 5.142
kJ/s
,s
2.177 2.957 3.091 4.203 1.103 1.492 1.313 1.837 0.975 1.535 1.048 2.125 0.947 1.508 1.287 1.703 0.993 1.452 1.345 1.837
kJ/s
,tR
0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
kJ/s
%?O
0.026 0.048 0.034 0.075 0.027 0.045 0.033 0.056 0.021 0.048 0.033 0.062 0.029 0.052 0.039 0.068 0.025 0.045 0.038 0.058
kJ/s
%©
0.865 1.282 1.058 1.452 0.865 1.403 1.067 1.559 0.924 1.317 1.013 1.535 0.876 1.379 1.096 1.529 0.822 1.262 0.971 1.449
kJ/s 1.152 1.050 0.783 1.456 0.977 1.463 1.198 1.431 0.726 1.189 0.772 1.103 1.142 1.388 1.403 1.256 0.767 1.182 0.785 1.802
=>= kJ/s 4.018 5.542 4.843 6.91 2.701 4.06 3.35 5.289 2.896 4.288 3.338 4.709 2.581 4.015 3.086 4.794 2.617 3.826 3.202 4.51
I,t kJ/s
I,
0.003 0.010 0.009 0.013 0.001 0.003 0.002 0.001 0.003 0.001 0.002 0.003 0.001 0.001 0.001 0.001 0.003 0.004 0.002 0.004
kJ/s
I,m
0.008 0.022 0.033 0.031 0.078 0.115 0.130 0.231 0.148 0.142 0.191 0.200 0.091 0.109 0.129 0.166 0.117 0.162 0.141 0.236
kJ/s
I,q
0.186 0.360 0.289 0.521 0.237 0.446 0.347 0.611 0.303 0.459 0.370 0.570 0.263 0.434 0.352 0.534 0.270 0.481 0.334 0.639
kJ/s
I,e
0.131 0.254 0.216 0.350 0.168 0.332 0.278 0.463 0.254 0.333 0.309 0.403 0.193 0.318 0.288 0.399 0.213 0.378 0.271 0.479
kJ/s
,s
0.100 0.209 0.186 0.307 0.062 0.125 0.095 0.163 0.082 0.128 0.079 0.165 0.074 0.117 0.094 0.150 0.060 0.125 0.084 0.159
kJ/s
,tR
0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000
kJ/s
?O
0.001 0.002 0.001 0.004 0.001 0.002 0.001 0.003 0.001 0.002 0.001 0.004 0.001 0.003 0.002 0.004 0.001 0.002 0.002 0.003
kJ/s
©
0.053 0.084 0.068 0.134 0.054 0.099 0.075 0.104 0.037 0.096 0.060 0.108 0.061 0.110 0.081 0.135 0.048 0.082 0.068 0.100
kJ/s
\?n
0.095 0.115 0.068 0.169 0.078 0.163 0.104 0.158 0.063 0.135 0.074 0.124 0.091 0.158 0.120 0.145 0.069 0.143 0.073 0.210
kJ/s 3.772 5.142 4.529 6.309 2.507 3.674 3.077 4.862 2.716 3.928 3.126 4.311 2.355 3.628 2.79 4.361 2.442 3.478 2.977 4.042
430

431

23
432

0.7
Base operation mode
A
Output warm exergy rate of drying

0.6 Scheme I
Scheme II
0.5 Scheme III
Scheme IV
cabinet (kJ/s)

0.4

0.3

0.2

0.1

0
0 20 40 60 80 100 120
Drying time (min)
0.7
Base operation mode
B
Output wet exergy rate of drying

0.6 Scheme I
Scheme II
0.5 Scheme III
Scheme IV
cabinet (kJ/s)

0.4

0.3

0.2

0.1

0
0 20 40 60 80 100 120
Drying time (min)
0.7
Base operation mode
C
Output total exergy rate of drying

0.6 Scheme I
Scheme II
0.5 Scheme III
Scheme IV
cabinet (kJ/s)

0.4

0.3

0.2

0.1

0
0 20 40 60 80 100 120
Drying time (min)
433 Figure 9. Variations in the output warm (A), wet (B), and total (C) exergy rates of drying
434 cabinet at T=70 ºC and Q=450 m3/h under different control schemes.
435

436 Figure 10 displays the variations in the output warm, wet, and total exergy rates of

24
437 drying cabinet at different experimental conditions under the control scheme III. The

438 input warm exergy rate of drying cabinet was enhanced by augmenting both temperature

439 and volumetric flow rate of drying air which in turn significantly boosted (p < 0.05) the

440 output warm exergy rate of drying cabinet (Figure 10A). There was no significant

441 difference (p > 0.05) in the output wet exergy of drying cabinet at different experimental

442 conditions (Figure 10B). In fact, the wet exergy rate of drying cabinet was remained

443 almost unchanged over time. This could be related to a slight accumulation of moisture

444 within drying medium owing to the exhaust air recirculation. Elevating both temperature

445 and volumetric flow rate of drying air significantly increased (p < 0.05) the total output

446 exergy rate of drying cabinet (Figure 10C). Notably, the total output exergy rate of drying

447 cabinet was significantly influenced by the warm exergy rate compared with the wet

448 exergy rate.

449

0.7
T=55℃, Q=360 m³/h
A
Output warm exergy rate of drying

0.6 T=55℃, Q=450 m³/h


0.5 T=70℃, Q=360 m³/h
T=70℃, Q=450 m³/h
cabinet (kJ/s)

0.4

0.3

0.2

0.1

0
0 30 60 90 120 150 180 210
Drying time (min)

25
0.7
T=55℃, Q=360 m³/h
Output wet exergy rate of drying 0.6
B
T=55℃, Q=450 m³/h
0.5 T=70℃, Q=360 m³/h
cabinet (kJ/s)

T=70℃, Q=450 m³/h


0.4

0.3

0.2

0.1

0
0 30 60 90 120 150 180 210
Drying time (min)
0.7
T=55℃, Q=360 m³/h
C
Output total exergy rate of drying

0.6 T=55℃, Q=450 m³/h


0.5 T=70℃, Q=360 m³/h
cabinet (kJ/s)

T=70℃, Q=450 m³/h


0.4

0.3

0.2

0.1

0
0 30 60 90 120 150 180 210
Drying time (min)
450 Figure 10. Variations in the output warm (A), wet (B), and total (C) exergy rates of
451 drying cabinet at different experimental conditions under the control scheme III.
452
453
454 Figure 11 represents the variations in the exergetic effectiveness of drying process

455 at T=70 ºC and Q=450 m3/h under different control schemes. The variations in the

456 exergetic effectiveness of drying process at different experimental conditions under the

457 control scheme III are also shown in Figure 12. The fluctuations observed in the exergetic

458 effectiveness of drying process under the developed control schemes could be related to

459 the real-time adjustment of exhaust air recirculation ratio. Unlike the controlled exhaust

460 air recirculation conditions, the exergetic effectiveness of drying process under the

461 base operation mode had a smooth shape. The exergetic effectiveness of drying process

462 was rapidly increased at the initial stage of drying process and then tended to decrease

26
463 towards the end of drying process. Increasing the rate of moisture removal from wood

464 chips at the inception of the process could be regarded as the main reason for the initial

465 increase of this parameter. However, the exergetic effectiveness of drying process was

466 diminished towards the end of the drying process because of a reduction in the rate of

467 moisture removal from wood chips. Interestingly, the exergetic effectiveness of drying

468 process was significantly improved (p < 0.05) by implementing all the developed control

469 schemes, particularly at the end of drying process (Figure 11). This occurred because of a

470 remarkable decrease in the energy consumption rate of drying system under controlled

471 exhaust air recirculation conditions as shown in Figure 16. Among the developed control

472 schemes, the highest exergetic effectiveness values were obtained in the control scheme

473 III while the lowest values observed in the control scheme IV (Figure 11). In the spite of

474 the fact that increasing both temperature and volumetric flow rate of drying air

475 markedly elevated the energy consumption rate, the exergetic effectiveness of drying

476 process was significantly improved (p < 0.05) because of an intensification in heat and

477 mass transfer coefficients at higher temperatures and volumetric flow rates of drying air.

478 As a matter of fact, higher temperatures and volumetric flow rates of drying air induced

479 more efficient utilization of the exergy supplied to the drying cabinet (Figure 12).

480
6
Base opeation mode
Scheme I
Exergetic effectiveness (%)

5 Scheme II
Scheme III
4 Scheme IV

0
0 20 40 60 80 100 120
481 Drying time (min)

27
482 Figure 11. Variations in the exergetic effectiveness of drying process at T=70 ºC and
483 Q=450 m3/h under different control schemes.
484
485
6
T=55℃, Q=360 m³/h
T=55℃, Q=450 m³/h
5
Exergetic effectiveness (%)

T=70℃, Q=360 m³/h


T=70℃, Q=450 m³/h
4

0
0 30 60 90 120 150 180 210
486 Drying time (min)
487 Figure 12. Variations in the exergetic effectiveness of drying process at different
488 experimental conditions under the control scheme III.
489
490
491 Figure 13 indicates the average exergetic effectiveness of drying process at different

492 experimental conditions under various control schemes. The exergetic effectiveness

493 improvement of the process compared with the base operation mode at different

494 experimental conditions under various control schemes is also exhibited in Figure 14. The

495 average exergetic effectiveness of drying process was found to be in the range of 1.4‒

496 3.0%. Coskun et al. (2009) reported the exergetic effectiveness of wood chips drying

497 process in an industrial-scale drum dryer as 4.4%. Ghasemkhani et al. (2016) determined

498 the exergetic effectiveness of apple slices drying process in a heat exchanger-equipped

499 rotating-tray convective dryer ranging from 0.4% to 2.2%. As a conclusion, the exergetic

500 effectiveness values obtained here agreed well with those reported previously.

501 Obviously, real-time exhaust air recirculation control could substantially improve (p <

502 0.05) the exergetic effectiveness of drying process when assessed in comparison with the

503 base operation mode (Figure 13). The control scheme III showed the highest exergetic

504 effectiveness values because of its good exhaust air recirculation controlling capability.

28
505 The highest exergetic effectiveness improvement of the process in the range of 46.5‒

506 75.9% was obtained when the control scheme III was implemented on drying system.

507 This value was in the range of 21.0‒46.1%, 19.6‒46.9%, and 11.5‒43.3% for the control

508 schemes I, II, and IV, respectively. Figure 15 portrays the exergy flow diagrams for the

509 developed control schemes under optimum operating conditions as well.

510

3.5
T=55℃, Q=360 m³/h
T=55℃, Q=450 m³/h
Average exergetic effectiveness (%)

3.0
T=70℃, Q=360 m³/h
T=70℃, Q=450 m³/h
2.5

2.0

1.5

1.0

0.5

0.0
Base operation Scheme I Scheme II Scheme III Scheme IV
mode
511
512 Figure 13. Average exergetic effectiveness of drying process at different experimental
513 trials under various control schemes.
514

80
T=55℃, Q=360 m³/h
T=55℃, Q=450 m³/h
Exergetic effectiveness improvement (%)

70
T=70℃, Q=360 m³/h
T=70℃, Q=450 m³/h
60

50

40

30

20

10

0
515 Scheme I Scheme II Scheme III Scheme IV

29
516 Figure 14. Exergetic effectiveness improvement of the process compared with the base
517 operation mode at different experimental trials under various control schemes.
518
519

520

30
521 Figure 15. Exergy flow diagrams for the developed control schemes under optimum
522 operating conditions.
523
524

525 Figure 16 demonstrates the variations in the energy consumption rate of the dryer at

526 T=70 ºC and Q=450 m3/h under different control schemes. The variations in the energy

527 consumption rate of the dryer at different experimental trials under the control scheme III

528 are also indicated in Figure 17. The controlled exhaust air recirculation conditions did not

529 significantly decrease (p > 0.05) the energy consumption rate of the dryer at the inception

530 of drying process compared with the base operation mode. This could be explained by the

531 high moisture content of the product and, subsequent, the low exhaust recirculation ratio

532 at the beginning of the process. However, the energy consumption rate of the dryer was

533 significantly decreased (p < 0.05) under the developed control schemes towards the end

534 of the process in comparison with the base operation mode. This reduction corresponded

535 with the reduction of the moisture content of wood chips as drying process progressed. In

536 better words, the developed control schemes lowered the energy consumption rate of

537 drying system at the end of drying process by recovering the exergy of exhaust air. This

538 reduction was more prominent under the control scheme III compared with the other

539 control schemes because of its flexible structure, making possible the efficient reuse of

540 the output warm exergy of drying cabinet. Although decreasing temperature and

31
541 volumetric flow rate of drying air decreased the rate of energy consumption at the

542 beginning of the process (Figure 17), increasing drying time negatively neutralized the

543 positive effect observed on the energy consumption rate of the dryer. In addition, both

544 temperature and volumetric flow rate of drying air did significantly affect (p > 0.05) the

545 energy consumption rate of drying system under the control scheme III at the end of the

546 process (Figure 17). This could be attributed to the efficient recirculation and reuse of the

547 exergetic content of exhaust air when the majority of the moisture content of the product

548 was removed.

549

8
Energy consumption rate of drying

7
6
system (kJ/s)

5
4
3
Base operation mode
2 Scheme I
Scheme II
1 Scheme III
Scheme IV
0
0 20 40 60 80 100 120
550 Drying time (min)
551 Figure 16. Variations in the energy consumption rate of the dryer at T=70 ºC and Q=450
552 m3/h under different control schemes.
553

554

32
8
T=55℃, Q=360 m³/h
Energy consumption rate of drying 7 T=55℃, Q=450 m³/h
6 T=70℃, Q=360 m³/h
T=70℃, Q=450 m³/h
system (kJ/s)

5
4
3
2
1
0
0 30 60 90 120 150 180 210
555 Drying time (min)
556 Figure 17. Variations in the energy consumption rate of the dryer at different
557 experimental trials under the control scheme III.
558

559 Figure 18 demonstrates the variations in the exergy destruction rate of drying system at

560 T=70 ºC and Q=450 m3/h under different control schemes. The variations in the exergy

561 destruction rate of drying system at different experimental trials under the control scheme

562 III are also indicated in Figure 19. Clearly, the exergy destruction rate of drying systems

563 was not significantly influenced (p > 0.05) at the inception of the drying process by the

564 developed control schemes. However, the exergy destruction rate of drying system was

565 significantly decreased (p < 0.05) by the developed control schemes towards the end of

566 the drying process in comparison with the base operation mode. In better words, the

567 developed control schemes could lower the exergy destruction rate of drying system at

568 the end of drying process by recovering the majority of the exergy of exhaust air. This

569 reduction was more marked under the control scheme III compared with the other control

570 schemes because of its flexible structure, making possible the efficient reuse of the output

571 warm exergy leaving the drying cabinet. According to Figure 19, the exergy destruction

572 rate of drying system was increased (p < 0.05) by increasing drying air temperature and

573 volumetric flow rate at the beginning of the process. However, there was no significant

574 difference (p > 0.05) in the rate of exergy destruction rate towards the end of the process

33
575 by increasing drying air temperature and volumetric flow rate.

576

7
Exergy destruction rate (kJ/s)

3
Base operation mode
2 Scheme I
Scheme II
1 Scheme III
Scheme IV
0
0 20 40 60 80 100 120
Drying time (min)
577
578 Figure 18. Variations in the exergy destruction rate of drying system at T=70 ºC and
579 Q=450 m3/h under different control schemes.
580
581
7
T=55℃, Q=360 m³/h
6 T=55℃, Q=450 m³/h
Exergy destruction rate (kJ/s)

T=70℃, Q=360 m³/h


5 T=70℃, Q=450 m³/h
4

0
0 30 60 90 120 150 180 210
582 Drying time (min)
583 Figure 19. Variations in the exergy destruction rate of drying system at different
584 experimental trials under the control scheme III.
585
586
587 Figure 20 exhibits the cumulative exergy destruction of drying system at different

588 experimental trials under the developed control schemes. Obviously, implementing all the

589 developed control schemes could significantly reduce (p < 0.05) the cumulative exergy

590 destruction of drying system. In addition, the control scheme III showed the

591 lowest cumulative energy consumption of drying system among the control

34
592 schemes applied. However, the effects of drying air temperature and volume flow rate on

593 the cumulative exergy destruction of drying system was insignificant (p > 0.05).

594
595
50
T=55℃, Q=360 m³/h
45 T=55℃, Q=450 m³/h
Cumulative exergy destruction (MJ)

40 T=70℃, Q=360 m³/h


T=70℃, Q=450 m³/h
35
30
25
20
15
10
5
0
Base operation Scheme I Scheme II Scheme III Scheme IV
596 mode

597 Figure 20. Cumulative exergy destruction of drying system at different experimental trials
598 under various control schemes.
599
600

601 Figures 21 and 22 respectively exhibit the cumulative energy consumption of the

602 dryer and the energy saving of the process at different experimental trials under the

603 developed control schemes. Obviously, implementing all the developed control schemes

604 could significantly reduce (p < 0.05) the cumulative energy consumption of the dryer.

605 The control scheme III showed the lowest cumulative energy consumption of the dryer

606 among the exhaust air recirculation control schemes applied. However, there was no clear

607 trend (p > 0.05) in the cumulative energy consumption of the dryer among different

608 experimental trials. According to Figure 22, the energy saving of the process under the

609 control scheme III was significantly higher (p < 0.05) than those of the other

610 control schemes. More specifically, the highest energy saving of the process in the range

35
611 of 30.7‒34.5% was obtained under the control scheme III. This value was in the range of

612 22.2‒30.0%, 26.0‒34.1%, and 20.9‒29.7% for the control schemes I, II, and IV,

613 respectively. Similar to the cumulative energy consumption of the dryer, the energy

614 saving of the process was not significantly influenced (p > 0.05) by experimental trials.

615 Taking into consideration the exergetic effectiveness improvement and energy saving of

616 the process, the control scheme III appeared to be a suitable approach for recovering the

617 majority of exergy leaving drying cabinet. It should be emphasized that the overall drying

618 time did not significantly change (p > 0.05) under the developed control schemes,

619 particularly control scheme III as shown in Figure 6. This occurred since the moisture

620 accumulation within the drying medium as the main reason for prolonging drying process

621 was avoided in the control scheme III.

60
T=55℃, Q=360 m³/h
Cumulative energy consumption of drying

T=55℃, Q=450 m³/h


50 T=70℃, Q=360 m³/h
T=70℃, Q=450 m³/h
40
system (MJ)

30

20

10

0
Base operation mode Scheme I Scheme II Scheme III Scheme IV
622
623 Figure 21. Cumulative energy consumption of the dryer at different experimental trials
624 under various control schemes.
625

36
50
T=55℃, Q=360 m³/h
45 T=55℃, Q=450 m³/h
Energy saving of drying process (%) T=70℃, Q=360 m³/h
40
T=70℃, Q=450 m³/h
35

30

25

20

15

10

0
Scheme I Scheme II Scheme III Scheme IV
626
627 Figure 22. Energy saving of the process at different experimental trials under various
628 control schemes.
629

630 Velić et al. (2003) found that a maximum energy saving of 14% could be theoretically

631 obtained in a typical spray drying process through exhaust air recirculation. The

632 maximum energy saving value (34.5%) obtained in this study was markedly higher than

633 that of Velić et al. (2003). Tippayawong et al. (2008) reported that liquefied natural gas

634 substitution with wood, exhaust air recirculation, thermal insulation improvement, and

635 temperature and humidity control could experimentally increase the thermal efficiency of

636 a longan dryer from 29% to 35%. In a similar study, Tippayawong et al. (2009) found that

637 the thermal efficiency of a longan fruit drying system could be increased from 19%

638 to 29% through changing heat transfer mode, recirculating exhaust air, improving

639 thermal insulation, and controlling temperature and humidity according to the

640 experimental results. The energy saving values obtained under the selected control

641 scheme (30.7‒34.5%) were in line with those of Tippayawong et al. (2009, 2008).

642 However, it should be noted that Tippayawong et al. (2009, 2008) achieved these values

643 through several improvements while only exhaust air recirculation was applied in the

37
644 present study. Golman and Julklang (2014) deduced that a significant energy saving up to

645 about 70% could theoretically be obtained for a spray dryer with a high ratio of exhaust

646 air recirculation. Ziegler et al. (2016) concluded in a simulation study that partial exhaust

647 air recirculation in a batch‐type dryer could potentially save up to 75% of the required

648 energy of the process compared with the base operation mode. Despite the fact that the

649 energy saving values obtained herein were lower than the claimed amounts in Golman

650 and Julklang (2014) and Ziegler et al. (2016), those values have not yet been

651 experimentally validated and realized. It should also be noted that a rigorous comparison

652 of the energy saving values reported in different papers could not likely be achieved

653 because of fundamental differences in the processes investigated, variables considered,

654 and assumptions made.

655

656 4. Concluding remarks and future trends

657 Three new real-time exergy-based control frameworks were introduced and

658 experimented for adjusting the recirculation ratio of exhaust air in a convective dryer. The

659 first control scheme adjusted the recirculation ratio of exhaust air by measuring the

660 output–input warm exergy ratio of drying cabinet. In addition to the output–input warm

661 exergy ratio of drying cabinet, a fixed and varying upper limit for the output wet exergy

662 rate of drying cabinet were respectively considered in the second and third control

663 schemes. The fourth scheme recirculated the exhaust air in proportion to the time of

664 drying process in a continuous manner. The exergetic effectiveness improvement and

665 energy saving of the process achieved using these control schemes were assessed with

666 respect to the base operation mode for decision-making on the most efficient control

667 schemes. As a case study, poplar wood chips were dried at two temperatures (55 and 70

668 ºC) and two volumetric flow rates (360 and 450 m3/h) of drying air. The following

38
669 concluding remarks could be derived from the present investigation:

670 • There was not significant change in drying time of poplar wood chips under

671 various control schemes.

672 • The third control scheme showed more fluctuations in the position of recirculation

673 flap over time.

674 • The developed control schemes did not significantly affect the output

675 warm exergy rate of drying cabinet compared with the base operation mode.

676 • The output wet exergy rate and the total output exergy rate of drying cabinet were

677 significantly higher for all the developed control schemes in comparison with the

678 base operation mode.

679 • Elevating both temperature and volumetric flow rate of drying air increased

680 the output warm exergy rate and the total output exergy rate of drying cabinet.

681 • There was no significant difference in the output wet exergy of drying cabinet at

682 different temperatures and volumetric flow rates of drying air.

683 • The average exergetic effectiveness of drying process was found to be in the range

684 of 1.4‒3.0%.

685 • The exergetic effectiveness of drying process was significantly improved by

686 implementing all the developed control schemes.

687 • The highest exergetic effectiveness improvement of the process in the range of

688 46.5‒75.9% was obtained under the third control scheme.

689 • Implementing all the developed control schemes could significantly reduce the

690 cumulative energy consumption of the dryer.

691 • The highest energy saving of the process in the range of 30.7‒34.5% was obtained

692 under the third control scheme.

39
693 • Taking into consideration the exergetic effectiveness improvement and energy

694 saving of the process, the third control scheme appeared to be a suitable approach

695 for recovering the majority of exergy leaving drying cabinet.

696 In general, the selected real-time exergy-based control framework appears to be a

697 practical and powerful tool for adjusting the recirculation ratio of exhaust air in drying

698 industry. However, economic and environmental aspects of such an implementation

699 should be assessed in order to make comprehensive decisions on the exhaust air

700 recirculation ratio. Accordingly, future research work should be directed towards

701 developing and examining real-time exergoeconomic- and exergoenvironmental-based

702 control frameworks for adjusting the recirculation ratio of exhaust air. Such advanced

703 analyses would further improve the efficiency, productivity, and sustainability of exhaust

704 air recirculation in drying technology.

705

706 Acknowledgments

707 The authors would like to express their appreciation to the Iranian Ministry of Science,

708 Research and Technology and the Leibniz Institute for Agricultural Engineering and

709 Bioeconomy (ATB) (Potsdam, Germany) for providing financial and technical support.

710

711 References

712 Aghbashlo, M., Kianmehr, M.H., Arabhosseini, A., 2009. Performance analysis of drying

713 of carrot slices in a semi-industrial continuous band dryer. J. Food Eng. 91, 99–108.

714 Aghbashlo, M., Kianmehr, M.H., Arabhosseini, A., 2008. Energy and exergy analyses of

715 thin-layer drying of potato slices in a semi-industrial continuous band dryer. Dry.

716 Technol. 26, 1501–1508.

40
717 Aghbashlo, M., Mandegari, M., Tabatabaei, M., Farzad, S., Mojarab Soufiyan, M.,

718 Görgens, J.F., 2018a. Exergy analysis of a lignocellulosic-based biorefinery annexed

719 to a sugarcane mill for simultaneous lactic acid and electricity production. Energy

720 149, 623–638.

721 Aghbashlo, M., Rosen, M.A., 2018. Exergoeconoenvironmental analysis as a new

722 concept for developing thermodynamically, economically, and environmentally

723 sound energy conversion systems. J. Clean. Prod. 187, 190–204.

724 Aghbashlo, M., Tabatabaei, M., Hosseini, S.S., Dashti, B.B., Mojarab Soufiyan, M.,

725 2018b. Performance assessment of a wind power plant using standard exergy and

726 extended exergy accounting (EEA) approaches. J. Clean. Prod. 171, 127–136.

727 Aghbashlo, M., Tabatabaei, M., Mohammadi, P., Khoshnevisan, B., Rajaeifar, M.A.,

728 Pakzad, M., 2017. Neat diesel beats waste-oriented biodiesel from the

729 exergoeconomic and exergoenvironmental point of views. Energy Convers. Manag.

730 148, 1–15.

731 Aghbashlo, M., Tabatabaei, M., Rastegari, H., Ghaziaskar, H.S., Valijanian, E., 2018c.

732 Exergy-based optimization of a continuous reactor applied to produce value-added

733 chemicals from glycerol through esterification with acetic acid. Energy 150, 351–

734 362.

735 Aghbashlo, M., Tabatabaei, M., Soltanian, S., Ghanavati, H., Dadak, A., 2019.

736 Comprehensive exergoeconomic analysis of a municipal solid waste digestion plant

737 equipped with a biogas genset. Waste Manag. 87, 485–498.

738 Brooker, D.B., Bakker-Arkema, F.W., Hall, C.W., 1992. Drying and storage of grain and

739 oilseeds. Westport: AVI. Springer Science & Business Media.

740 Colak, N., Erbay, Z., Hepbasli, A., 2013. Performance assessment and optimization of

741 industrial pasta drying. Int. J. Energy Res. 37, 913–922.

41
742 Colak, N., Hepbasli, A., 2007. Performance analysis of drying of green olive in a tray

743 dryer. J. Food Eng. 80, 1188–1193.

744 Coskun, C., Bayraktar, M., Oktay, Z., Dincer, I., 2009. Energy and exergy analyses of an

745 industrial wood chips drying process. Int. J. Low-Carbon Technol. 4, 224–229.

746 Dincer, I., Rosen, M.A., 2005. Thermodynamic aspects of renewables and sustainable

747 development. Renew. Sustain. Energy Rev. 9, 169–189.

748 Erbay, Z., Icier, F., 2009. Optimization of hot air drying of olive leaves using response

749 surface methodology. J. Food Eng. 91, 533–541.

750 Erbay, Z., Koca, N., 2012. Energetic, exergetic, and exergoeconomic analyses of spray-

751 drying process during white cheese powder production. Dry. Technol. 30, 435–444.

752 Gasparatos, A., El-Haram, M., Horner, M., 2008. A critical review of reductionist

753 approaches for assessing the progress towards sustainability. Environ. Impact

754 Assess. Rev. 28, 286–311.

755 Genc, S., Hepbasli, A., 2015. Performance assessment of a potato crisp frying process.

756 Dry. Technol. 33, 865–875.

757 Ghasemkhani, H., Keyhani, A., Aghbashlo, M., Rafiee, S., Mujumdar, A.S., 2016.

758 Improving exergetic performance parameters of a rotating-tray air dryer via a simple

759 heat exchanger. Appl. Therm. Eng. 94, 13–23.

760 Golman, B., Julklang, W., 2014. Analysis of heat recovery from a spray dryer by

761 recirculation of exhaust air. Energy Convers. Manag. 88, 641–649.

762 Holman, J.P., 2001. Analysis of experimental data, in: Experimental Methods for

763 Engineers. McGraw Hill, Singapore, pp. 48–143.

764 Karthikeyan, A.K., Murugavelh, S., 2018. Thin layer drying kinetics and exergy analysis

765 of turmeric (Curcuma longa) in a mixed mode forced convection solar tunnel dryer.

766 Renew. energy 128, 305–312.

42
767 Khanali, M., Aghbashlo, M., Rafiee, S., Jafari, A., 2013. Exergetic performance

768 assessment of plug flow fluidised bed drying process of rough rice. Int. J. Exergy 13,

769 387–408.

770 Mandegari, M.A., Pahlavanzadeh, H., 2013. A Study on the Optimization of an Air

771 Dehumidification Desiccant System. J. Therm. Sci. Eng. Appl. 5, 41002.

772 Martin, H., Gnielinski, V., Mewes, D., Steiner, D., Stephan, K., Schaber, K., Vortmeyer,

773 D., Kabelac, S., 2002. VDI Wärmeatlas: Berechnungsblätter für den

774 Wärmeübergang. Verein Dtsch. Ingenieure.

775 Mkwananzi, T., Mandegari, M., Görgens, J.F., 2019. Disturbance modelling through

776 steady-state value deviations: The determination of suitable energy indicators and

777 parameters for energy consumption monitoring in a typical sugar mill. Energy 176,

778 211–223.

779 Mojarab Soufiyan, M., Dadak, A., Hosseini, S.S., Nasiri, F., Dowlati, M., Tahmasebi, M.,

780 Aghbashlo, M., 2016. Comprehensive exergy analysis of a commercial tomato paste

781 plant with a double-effect evaporator. Energy 111, 910–922.

782 Nazghelichi, T., Kianmehr, M.H., Aghbashlo, M., 2010. Thermodynamic analysis of

783 fluidized bed drying of carrot cubes. Energy 35, 4679–4684.

784 Perry, R.H., Green, D.W., Maloney, J.O., 2008. Perry’s Chemical Engineers’ Handbook,

785 McGraw and Hill. New York, USA.

786 Ranjbaran, M., Emadi, B., Zare, D., 2014. CFD simulation of deep-bed paddy drying

787 process and performance. Dry. Technol. 32, 919–934.

788 Sarker, M.S.H., Ibrahim, M.N., Aziz, N.A., Punan, M.S., 2015. Energy and exergy

789 analysis of industrial fluidized bed drying of paddy. Energy 84, 131–138.

790 Şevik, S., Aktaş, M., Dolgun, E.C., Arslan, E., Tuncer, A.D., 2019. Performance analysis

791 of solar and solar-infrared dryer of mint and apple slices using energy-exergy

43
792 methodology. Sol. Energy 180, 537–549.

793 Sheikhshoaei, H., Dowlati, M., Aghbashlo, M., Rosen, M.A., 2019. Exergy analysis of a

794 pistachio roasting system. Dry. Technol. 1–19.

795 https://doi.org/10.1080/07373937.2019.1649276

796 Shukuya, M., 2012. Exergy: Theory and Applications in the Built Environment. Springer

797 Science & Business Media.

798 Tippayawong, N., Tantakitti, C., Thavornun, S., 2008. Energy efficiency improvements in

799 longan drying practice. Energy 33, 1137–1143.

800 Tippayawong, N., Tantakitti, C., Thavornun, S., Peerawanitkul, V., 2009. Energy

801 conservation in drying of peeled longan by forced convection and hot air

802 recirculation. Biosyst. Eng. 104, 199–204.

803 Tzempelikos, D.A., Vouros, A.P., Bardakas, A. V, Filios, A.E., Margaris, D.P., 2014.

804 Case studies on the effect of the air drying conditions on the convective drying of

805 quinces. Case Stud. Therm. Eng. 3, 79–85.

806 Velić, D., Bilić, M., Tomas, S., Planinić, M., 2003. Simulation, calculation and

807 possibilities of energy saving in spray drying process. Appl. Therm. Eng. 23, 2119–

808 2131.

809 Yang, A., Jin, S., Shen, W., Cui, P., Chien, I.-L., Ren, J., 2019a. Investigation of energy-

810 saving azeotropic dividing wall column to achieve cleaner production via heat

811 exchanger network and heat pump technique. J. Clean. Prod.

812 Yang, A., Zou, H., Chien, I.-L., Wang, D., Wei, S., Ren, J., Shen, W., 2019b. Optimal

813 design and effective control of triple-column extractive distillation for separating

814 ethyl acetate/ethanol/water with multiazeotrope. Ind. Eng. Chem. Res. 58, 7265–

815 7283.

816 Yildirim, N., Genc, S., 2017. Energy and exergy analysis of a milk powder production

44
817 system. Energy Convers. Manag. 149, 698–705.

818 Yildirim, N., Genc, S., 2015. Thermodynamic analysis of a milk pasteurization process

819 assisted by geothermal energy. Energy 90, 987–996.

820 Ziegler, T., Jubaer, H., Schütz, M., 2016. Increasing the Energy Efficiency of Batch-Type

821 Drying with Partial Air Recirculation. Chemie Ing. Tech. 88, 208–214.

822 Zohrabi, S., Seiiedlou, S.S., Aghbashlo, M., Scaar, H., Mellmann, J., 2019. Enhancing the

823 exergetic performance of a pilot-scale convective dryer by exhaust air recirculation.

824 Dry. Technol. 1–16.

825

45
Research highlights:

- Exhaust air recirculation of a convective air dryer was exergetically controlled.

- The developed control schemes improved the exergetic effectiveness of the process.

- The optimal control method improved the exergetic effectiveness of the process up to 75%.

- The selected control technique reduced the overall energy needs of the process up to 34%.

1
Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

Paper title: Energy saving in a convective dryer by using novel real-time exergy-based control schemes
adjusting exhaust air recirculation

Authors: Saman Zohrabi, Mortaza Aghbashlo, Seyed Sadegh Seiiedlou, Holger Scaar, Jochen Mellmann

Das könnte Ihnen auch gefallen