You are on page 1of 10

Solar Energy 174 (2018) 73–82

Contents lists available at ScienceDirect

Solar Energy
journal homepage: www.elsevier.com/locate/solener

Time-dependent modelling of nanofluid-based direct absorption parabolic T


trough solar collectors

G.J. O’Keeffea, , S.L. Mitchella, T.G. Myersb, V. Cregana
a
Department of Mathematics and Statistics, University of Limerick, Co. Limerick, Ireland
b
Centre de Recerca Matemàtica, Campus UAB Edifici C, 08193 Bellaterra, Barcelona, Spain

A R T I C LE I N FO A B S T R A C T

Keywords: In this paper we propose a time-dependent, three-dimensional model for the efficiency of a nanofluid-based
Nanofluid direct-absorption parabolic trough solar collector under a turbulent flow regime. The model consists of a system
Direct absorption solar collector of equations: a partial differential equation for conservation of energy, and a time-dependent radiative transport
Parabolic trough equation describing the propagation of solar radiation through the nanofluid. Writing the model in di-
Solar energy
mensionless form reveals four controlling dimensionless numbers: one describing the relative importance of
conduction and advection and three describing the heat loss to the surroundings. Realistic parameter values are
applied to reduce the model further and these indicate that two of the dimensionless groups have a much smaller
impact on the performance of the solar collector. We use the resulting solution for the temperature to calculate
an analytic expression for the collector’s efficiency. This expression permits optimisation of design parameters
such as particle loading, incoming radiative intensity, receiver dimensions, the inlet temperature, and solar
concentration ratio.

1. Introduction collector. Li et al. (2016) show that by focusing incoming radiation, it is


possible for the surface of a DAPSC’s receiver to experience lower
Global capacity for generating concentrating solar thermal power temperatures than its center-line. Previous studies have compared the
(CSP) increased by more than 40% per year on average between 2008 efficiencies of direct-absorption and surface-based solar collectors and
and 2012, which placed CSP amongst the fastest growing forms of en- hypothesise that direct-absorption solar collectors’ lower surface tem-
ergy generation (Ellabban et al., 2014). There are multiple ways to peratures could make them more efficient (Taylor et al., 2011; Khullar
generate CSP, for example, Xie et al. (2011) numerically and experi- et al., 2012; Xu et al., 2015; Li et al., 2016; O’Keeffe et al., 2016,
mentally study a point focus solar collector using high concentration 2018a,b). However, conventional DAPSCs are limited because standard
Fresnel lens, however, parabolic trough systems have driven most of the working fluids are inefficient at absorbing sunlight; for example,
recent CSP capacity growth (Sawin et al., n.d.). Surface-based parabolic Otanicar et al. (2009) show that water only absorbs 13% of the avail-
trough solar collectors (SPSCs) are the predominant parabolic trough able solar energy at a depth of 1 cm. Therefore, SPSCs outperform
system design. SPSCs usually consist of a working fluid flowing through DAPSCs using standard working fluids. Theoretical and experimental
metallic pipes. These pipes heat up as they absorb incoming solar ra- studies show that nanofluids have enhanced optical properties for ab-
diation, and this heat is then absorbed by the working fluid. Black-body sorbing solar radiation over their base-fluids (Otanicar et al., 2009;
emissions from the SPSC surface are a major source of inefficiency in Taylor et al., 2011). A nanofluid is a colloidal suspension of nano-
this design since an SPSC’s surface is the hottest part of the collector (Li particles in a liquid medium. In a nanofluid, solar radiation is atte-
et al., 2016). Direct-absorbing parabolic trough solar collectors nuated much faster due to the nanoparticles absorbing and scattering
(DAPSCs) are an alternative (albeit less popular) form of parabolic the solar radiation that propagates through the receiver. This has led to
trough CSP production; in a DAPSC, the working fluid is heated volu- the development of nanofluid-based direct-absorption parabolic trough
metrically by incoming radiation rather than at the surface of the re- solar collectors (NDAPSCs).
ceiver. Martinopoulos et al. (2010) show experimentally that the effi- Several studies model NDAPSCs in an attempt to better understand
ciency of a translucent polycarbonate direct-absorbing solar collector and predict their performance. Khullar et al. (2012) consider a steady-
can be similar to that of low-cost flat plate commercially available state two-dimensional model for the temperature and efficiency of an


Corresponding author.
E-mail addresses: gary.okeeffe@ul.ie (G.J. O’Keeffe), sarah.mitchell@ul.ie (S.L. Mitchell), tmyers@crm.cat (T.G. Myers), vincent.cregan@ul.ie (V. Cregan).

https://doi.org/10.1016/j.solener.2018.08.073
Received 29 March 2018; Received in revised form 6 August 2018; Accepted 26 August 2018
Available online 05 September 2018
0038-092X/ © 2018 Elsevier Ltd. All rights reserved.
G.J. O’Keeffe et al. Solar Energy 174 (2018) 73–82

Nomenclature η efficiency [–]


CA solar concentration ratio [–]
R receiver radius [m] ϒT transmittance [–]
σ Stefan’s constant [kg s−1 K−4] ϒR reflectivity [–]
L receiver length [m] r∗ coordinate [m]
u mean fluid velocity [m s−1] x∗ coordinate [m]
T ∗ temperature [K] ϕ coordinate [rad]
Gs∗ incident radiative heat flux [W m−2]
ρ density [kg m−3] Subscripts
k thermal conductivity [Wm−1 K−1]
cp heat capacity [J kg−1 K−1] bf base fluid
fv nanofluid particle volume fraction [–] np nanoparticle
Gm, A, B, β0, β1 fitting parameters [–] nf nanofluid
Pe Peclet number [–] O outlet
Re Reynolds number [–] I inlet
γ , φ, τ dimensionless parameters [–] A ambient
∊ emissivity [–]

Al/Therminol® VP-1 NDAPSC subject to coupled radiative and diffusive absorption parabolic trough solar collector under a turbulent flow re-
heat transfer in an absorbing, emitting, and scattering medium under gime. In Section 2.2 the system’s conservation of energy is modelled.
plug flow. They compare a numerical treatment of their model with Time-dependence is introduced into this model via the source term, and
experimental data for conventional concentrating parabolic solar col- in Section 2.3 two potential source terms which operate on two dif-
lectors. Menbari et al. (2016) propose a steady-state model for a CuO/ ferent time-scales are described. The first of these terms represents the
Water NDAPSC subject to steady turbulent depth-dependent flow. They effect of dynamic cloud cover, and the second represents the effect of
validate the model by comparing a finite difference solution for the the Earth’s rotation about its axis. We rescale and non-dimensionalise
temperature with experimental results. Xu et al. (2015), while com- the model in Section 2.4 to yield five dimensionless controlling groups.
paring the performance of a medium-temperature (80–250 °C) NDAPSC These were also obtained by the authors in previous research (O’Keeffe
to that of an SPSC, show that the NDAPSC’s working fluid temperature et al., 2018b). Realistic parameter values applied to these groups de-
distribution is more uniform than that of the SPSC’s, and therefore, an monstrates that two of the dimensionless parameters have a compara-
NDASC can have greater efficiency than an SPSC within a preferred tively small impact on the model. In Section 2.6 we describe an analytic
working temperature range. O’Keeffe et al. (2018a,b) consider a steady- method for solving the governing system of equations. This method
state model for the temperature and efficiency of an Al/Therminol® VP- leads to an expression for the temperature of the nanofluid as it flows
1 NDAPSC. Unlike Menbari et al. (2016) and Khullar et al. (2012), through the collector. We use this analytic expression for the tem-
O’Keeffe et al. (2018b) obtain an analytic expression for the tempera- perature to evaluate the collector’s efficiency in Section 2.7 before
ture of the nanofluid as it flows through the NDAPSC which they used discussing the collector’s performance further in Section 3.
to calculate collector efficiency. A comprehensive review of the litera-
ture surrounding the application of heat-mirrors to NDAPSCs can be 2. Model
found in O’Keeffe et al. (2018b). A heat-mirror is a selectively trans-
missive/reflective material that is highly transparent at short wave- 2.1. Problem configuration
lengths, but highly reflective at long wavelengths, and was introduced
for use in solar-thermal energy conversion applications in the 1970s by The NDAPSC is modelled as a cylinder, wherein the variables, x ∗, r ∗,
Fan and Bachner (1976). Previous research had suggested that heat- and ϕ define a three-dimensional system such that x ∗ is the axial co-
mirror coatings could improve the efficiency of an NDAPSC (Taylor ordinate, r ∗ is the radial coordinate, and ϕ is the azimuthal angle (note
et al., 2011; Khullar et al., 2014; Li et al., 2016), however, O’Keeffe that ∗ denotes a dimensional variable). Fig. 1(b) shows the system
et al. (2018b) show that this is not always the case: for lower tem- geometry in more detail. The nanofluid enters the receiver at the inlet
peratures an uncoated system may be more efficient. Also, as the solar ( x ∗ = 0 ) at an inlet temperature of TI∗, before being pumped through the
concentration ratio increases, an uncoated NDAPSC becomes more ef- receiver. As the nanofluid flows towards the outlet ( x ∗ = L), it heats up
ficient than an NDAPSC coated with a heat-mirror; at higher inlet before exiting the system with temperature TO∗ .
temperatures, the concentration ratio required for an uncoated NDAPSC
to be more efficient than a coated NDAPSC increases. 2.2. Conservation of energy
Although several researchers have studied the performance of
NDAPSCs, there are still significant gaps in the literature. Most notably, The equation describing the conservation of heat energy is similar to
the solar intensity at a fixed position on Earth is constantly changing the system described in O’Keeffe et al. (2018b), however, that paper
(Kimball, 1935), however, existing NDAPSC models assume that in- only concerns the steady-state case. Here, conservation of energy in the
coming solar intensity is constant. Kolb (2011) notes how a solar col- system is given by
lector must, by its nature, operate under dynamic conditions. The pipes
in a solar collector expand as they are heated, and this expansion ρnf cp, nf [Tt∗∗ + u ·∇T ∗] = knf ∇2 T ∗ + q, (1)
process produces mechanical strains. A solar collector needs to with-
stand these temperature fluctuations over its life cycle and perform where u = (u∗, w∗, is the fluid velocity,
v ∗) T ∗ (x ∗, r ∗,
ϕ, is the fluid
t ∗)
efficiently under realistic operating conditions; therefore, one must be temperature, q (r ∗, ϕ, t ∗) is the time-dependent source term, and the
able to predict the relationship between fluctuating solar intensity and physical properties, ρnf , cp, nf , and knf represent, the nanofluid’s density,
solar collector temperature. specific heat capacity, and thermal conductivity respectively. The inlet
This paper proposes an approximate analytic expression for the temperature condition is T ∗|x∗= 0 = TI∗, whilst the initial condition is
temperature and efficiency of a time-dependent nanofluid-based direct- T ∗|t∗= 0 = T0∗. At the surface of the receiver there is no slip, and thus
u|r ∗= R = 0 . The radiative boundary condition at the surface of the

74
G.J. O’Keeffe et al. Solar Energy 174 (2018) 73–82

Fig. 1. (a) Cross section of NDAPSC with length L and receiver with radius R, and (b) receiver geometry.

receiver is ⎧0 if ϕ > ϕcrit


Gs∗,1 (ϕ, t ∗) = G∗ (t ∗)2f
σ ∊ ∗4 ∗4 ⎨ s R 1 − 1sinϕ if ϕ ⩽ ϕcrit ,
Tr∗∗ |r ∗= R = (T A −T |r ∗= R ), ⎩ (6)
knf (2)
and
where σ is Stefan’s constant, ∊ the emissivity constant, and T A∗ the
ambient temperature. Furthermore, symmetry in the system implies ⎧0 if ϕ < −ϕcrit
Gs∗,2 (ϕ, t ∗) = Gs (t ∗)2f 1
Tϕ∗ |ϕ =−π /2, π /2 = 0. (3) ⎨ R 1 − sin(−ϕ) if ϕ ⩾ −ϕcrit ,
⎩ (7)
The Reynolds number is a well-known ratio that determines the where Gs∗ (t ∗) is the solar intensity at the aperture, ϕcrit is the maximum
flow regime of a system, and is given by Re = 2u ∗R/ ν , where u ∗ is the angle of incoming reflected solar radiation, and f is the focal length of
mean velocity of the fluid, and ν is the kinematic viscosity of the the parabolic reflector. The critical angle, ϕcrit , is limited by the width of
working fluid. Following the case-study from O’Keeffe et al. (2018b), the reflector, W, and is defined by
this paper focuses on modelling a scenario where the Reynolds number
is 13542, therefore, the flow is turbulent. For large Reynolds numbers, (W /2)2−4f 2 ⎤
ϕcrit = arctan ⎡ .
the turbulent thermal diffusion is much stronger than the molecular ⎢
⎣ 2fW ⎥
⎦ (8)
thermal diffusion (Elperin et al., 1996). O’Keeffe et al. (2018b) note
Next we introduce the constant Π , the overall fraction of incoming
that since the flow is turbulent, the temperature in the receiver’s cross
radiation that is absorbed into the nanofluid, i.e.,
section is approximately constant, i.e., T ∗ (x ∗, r ∗, ϕ, t ∗) ≃ T ∗ (x ∗, t ∗) . We
reduce (1) following the simplification method outlined in O’Keeffe 1 π /2 R 1
et al. (2018b): the terms in (1) are integrated over the cross section to
Π=
πRCA ϒT2 ϒRGs∗ (t ∗)
∫−π /2 ∫0 r ∗q (r ∗, ϕ, t ∗)dr ∗dϕ = 1− β1
,
obtain the one-dimensional model ( β0
2
+1 )
(9)
R2 ∗ R2 ∗ ∗ knf 2
Tt∗ + u T x∗ = ⎛ R Tx∗∗x∗ + RT ∗∗ |r ∗= R ⎞
⎜ ⎟

2 2 ρnf cp, nf ⎝ 2 r where CA is the solar concentration ratio, ϒT is the transmissivity of the

glass envelop to solar radiation, and ϒR is the reflectivity of the para-
1 R π /2
+
ρnf cp, nf 0
∫ ∫
−π /2
r q (r , ϕ, t ∗)dr ∗dϕ.
∗ ∗ ∗
bolic reflector. This fraction is used to rewrite (4) yielding
(4)
R2 ∗ knf ⎛ R2 ∗ Rσ ∊ ∗4 ∗4 ⎞
We use the source term similar to that proposed in O’Keeffe et al. (Tt∗ + u ∗Tx∗∗) = ⎜ Tx∗x∗ + (T A −T ) ⎟
2 ρnf cp, nf ⎝ 2 knf ⎠
(2018b), but, here we include time-dependence
ΠR ϒT2 ϒRCA Gs∗ (t ∗)
+ .
⎛ ⎞ ρnf cp, nf (10)
β0 β1 Gs∗,1 (ϕ, t ∗) Gs∗,2 (ϕ, t ∗)
q (r ∗, ϕ, t ∗) = ⎜ + ⎟,
2r ∗ ⎜ β1+ 1 β1+ 1

⎝( β
1 + 2R0 (R−r ∗) ) ( β
1 + 2R0 (R + r ∗) )


2.3. Solar intensity
(5)
where β0 and β1 are the dimensionless fitting parameters which were In this section we propose two realistic time-dependent examples of
first introduced by Cregan and Myers (2015) to approximate the source the solar intensity at the aperture, Gs∗ (t ∗) . Scenario 1 models when a
term in a parallel-plate nanofluid-based direct absorption solar col- cloud passes over the receiver leading to a sharp decrease in solar in-
lector, and later adapted to model the source term in an NDAPSC in tensity, while Scenario 2 models the slower variation of solar intensity
O’Keeffe et al. (2018a,b). We note that these fitting parameters vary over the course of a day. Even though this paper primarily discusses
with: nanofluid particle volume fraction, receiver radius, type of na- these two scenarios, we emphasise that the solution method in Section
noparticle, and type of base-fluid. The 1/ r ∗ term in (5) describes the 2.6 is independent of Gs∗ (t ∗) , and so the model is easily extended to
concentration of incoming solar radiation as it gets closer to the para- incorporate alternative time-dependent representations of solar in-
bolic reflector’s focal line, the power-law terms are a result of the in- tensity.
coming and outgoing solar radiative intensity decaying as it gets ab- In Scenario 1, when a cloud covers the receiver, we presume an
sorbed into the nanofluid, while the ϕ -dependent functions are given by instantaneous drop in solar intensity. This is modelled as

75
G.J. O’Keeffe et al. Solar Energy 174 (2018) 73–82

Gs∗ (t ∗) = Gm∗ (1−0.5H(t ∗−tc∗)), (11) magnitudes of the dimensionless parameter values, 1/Pe, γ , φ , and τ ,
with a view towards simplifying the dimensionless conservation of
where Gm∗is the maximum solar intensity at the aperture, H(·) is the
energy equation, (15). Table 3 provides values for these dimensionless
Heaviside step function, and tc∗ is the time when the cloud shades the
parameters when TI∗ and fv are varied in the case study from Section
collector. In the case of Scenario 2 we use three weeks of minute by
2.5. From this table, we see that 1/Pe ranges between 3 × 10−8 and
minute data on incoming solar radiation beginning on 1st June 2015
5 × 10−8 while φ is O(10−2) . Consequently, we may neglect terms in-
from the University of Oregon’s Solar Radiation Monitoring Laboratory
cluding 1/Pe, and φ , and approximate (15) via
(University of Oregon Solar Radiation Monitoring Laboratory, 2015).
The mean solar irradiation at each time of the day is obtained by Tt + Tx = Gs (t ) + γ −τ 4. (17)
averaging across all 21 days in the database (dot-dashed grey line in
We note that neglecting 1/Pe and φ results in errors of the order 10−6%
Fig. 2). The data is approximated via
and 1% respectively.
w ∗− B 2 Since (17) is a first-order linear partial differential equation, it has
Gs∗ (t ∗) = Gm∗ e−A ( L t ), (12)
an analytic solution of the form:
where the parameters: Gm∗
= 849.4 W m−2 ,
A = 2.595 × and 10−5, t
B = 492.49 are obtained via a least-squares fitting routine in Matlab. T (x , t ) = ∫t−x Gs (s) ds + (γ−τ 4) x, (18)
The associated standard error score is 34.96 W m−2. Fig. 2 compares
the fitted function (black line) to the data (dot-dashed grey line). In which may be obtained via the method of characteristics. We remind
general, the experimental data and the fitted function match well: as the reader that this general solution can be applied to any time-de-
expected, the solar intensity gradually increases in the morning until pendent source term, however, in the specific context of Scenarios 1
midday before gradually decreasing for the rest of the day. and 2, (18) has the particular solutions
1 1
2.4. Dimensional analysis T (x , t ) = x − (t −tc )H(t −tc ) + (t −x −tc )H(t −x −tc ) + (γ −τ 4 ) x ,
2 2 (19)

We define the dimensionless variables or equivalently


4
x ∗ = lx , t∗ =
l
t, T ∗ = TI∗ + ΔT T , Gs∗ (t ∗) = Gm∗ Gs (t ), ⎧ (1 + γ −τ ) x if t < tc
u∗ (13) T (x , t ) = (1 + γ −τ 4 ) x −1/2(t −tc ) if tc ⩽ t ⩽ x + tc
⎨ 4
where ⎩ (1/2 + γ −τ ) x if t > x + tc , (20)
2ΠϒT2 ϒRCA Gm∗ L for Scenario 1, and
ΔT = ,
u ∗ρnf cp, nf R (14) π
T (x , t ) = [erf( A (B−t )) + erf( A (B + x −t ))] + (γ −τ 4 ) x ,
is chosen since the source term is driving the temperature variation. 2 A (21)
Non-dimensionalising (10) yields for Scenario 2 (where erf(·) is the error function). Supplementary
1 asymptotic analysis based on how quickly the solar intensity is chan-
Tt + Tx = Txx + γ + Gs (t )−(τ + φT ) 4 ,
Pe (15) ging (see Appendix A) shows that (21) is approximately equivalent to
the simpler analytic expression
where the dimensionless parameters are
ρnf cp, nf lu ∗ T (x , t ) = (Gs (t ) + γ −τ 4 ) x . (22)
σ ∊ T A∗4 γ1/4TI∗ γ1/4 ΔT
Pe = , γ= , τ= , φ= .
knf ΠϒT2 ϒRCA Gm∗ T A∗ T A∗ (16) Note that in Section 3 we sometimes refer to (22) rather than (21)
because (22) is simpler and easier to interpret. For example, it is ob-
These four dimensionless numbers also appear in the steady-state model
vious from (22) that when Gs (t ) > γ −τ 4 , then TO∗ > TI∗, and conversely
proposed in O’Keeffe et al. (2018b): The Peclet number, Pe, describes when Gs (t ) < γ −τ 4 , then TO∗ < TI∗ .
the ratio of advection to thermal diffusion, γ is the ratio of absorbed
background radiation to absorbed solar radiation, while τ and φ de-
2.7. Efficiency
scribe the relative magnitude of emitted black-body radiation. The di-
mensionless initial and inlet conditions are T|t = 0 = T|x = 0 = 0 .
Duffie and Beckman (2013) define the instantaneous collector effi-
ciency, ηt (t ∗) , as the ratio of usable thermal energy to incident solar
2.5. Case study
energy, i.e.,
As a case study for exploring our model further, we consider an
Aluminum/Therminol® VP-1 nanofluid and, unless otherwise stated, we
use the parameter values given in the case study described in O’Keeffe
et al. (2018b). That study compares the performance of an NDAPSC
coated with a heat-mirror to an NDAPSC not coated with a heat-mirror
using a steady-state model. In Section 3 we use the time-dependent
model to similarly compare the performance of these two NDAPSC
design variations. The parameter values used in this paper are stated in
Tables 1 and 2. Also, we calculate the nanofluid’s thermophysical and
optical properties via the appropriate formulas detailed in O’Keeffe
et al. (2018b).

2.6. Solution method


Fig. 2. Experimentally observed incoming radiative intensity (University of
Asymptotic analysis is a widely used approximation technique, see Oregon Solar Radiation Monitoring Laboratory, 2015) (dot-dashed grey line)
for example Veeraragavan et al. (2012), Cregan and Myers (2015), and and approximated incoming radiative intensity (black line) over the .course of a
O’Keeffe et al. (2016, 2018a,b). This section explores the relative day.

76
G.J. O’Keeffe et al. Solar Energy 174 (2018) 73–82

Table 1
Physical parameters used in case study where the subscripts np , bf , and nf re-
present nanoparticle, base-fluid, and nanofluid respectively. (Rakić, 1995;
Khullar et al., 2012; Giovannetti et al., 2014; O’Keeffe et al., 2018b; Solutia,
n.d.)
Symbol Value Units

R, L, W 0.035, 8, 5 m
CA 22.7364 –
ϒT 0.96 –
ϒR 0.93 –
∗ 20 °C
TA
σ 5.67e−8 kg s−1 K−4
∊ 0.92 –
Q 9.12e−4 m3s−1
ρbf 1083−0.91TI∗ + 7.8e−4TI∗2−2.37e−6TI∗3 kg m−3
ρnp 2700 kg m−3
kbf 0.14−8.2e−5TI∗−1.9e−7TI∗2 + 2.5e−11TI∗3−7.3e−15TI∗4 W m−1 K−1
knp 247 W m−1 K−1
cp, bf 1498 + 2.41TI∗ + 6e−3TI∗2−3e−5TI∗3 + 4.4e−8TI∗4 J kg−1 K−1 Fig. 3. Temperature along the length of an NDAPSC for fv = 0 (dashed line),
cp, np 900 J kg−1 K−1 fv = 0.0005 (dot-dashed line), and fv = 0.006 (solid line), where TI = 200 ° C and
β0, β1 0.5, 46.8 (when fv = 0.006 and R = 0.035 m ) – Gs∗ (t ∗) = Gs∗ = 1000 W/m2.

arbitrarily choose tc∗ = tm∗ /2 .


Table 2
NDAPSC optical parameter values used in case study, (Fan and Bachner, 1976;
Khullar et al., 2012; Giovannetti et al., 2014).
3. Results
Parabolic Tube/envelope Heat-mirror
mirror
This section uses a parameter space exploration based around the
Material Low-iron Low-iron Sn-doped In203 & values from Tables 1 and 2 to compare various aspects of a collector’s
glass antireflective glass Corning 7059 glass performance. Fig. 3 shows system temperatures along the length of an
Reflectivity 0.93 – 0.912 (Radiative heat NDAPSC with a constant solar intensity at the aperture (1000 W/m2).
loss)
Since this solar intensity is constant, there is no mechanism for time-
Transmittance – 0.96 0.90 (Incoming solar
radiation) dependence in this example and so the system is running at a steady-
state. Therefore, in Fig. 3, system temperatures are calculated via the
steady-state model proposed in O’Keeffe et al. (2018b). In this section
ρnf cp, nf u ∗πR2 (T ∗ (L, t )−TI∗ ) we use this figure (and, more generally, the steady state model from
ηt (t ∗) = . O’Keeffe et al. (2018b)) as a base-case scenario for highlighting the
Gs∗ (t ∗) LCA 2πR (23)
importance of the time-dependent results in this study.
However, this definition is not appropriate in the case of a time-de- Fig. 4 shows the piecewise temperature profile for Scenario 1 (i.e.,
pendent model because (23) merely offers a snapshot of the efficiency cloud cover), as calculated by (19) or (20). Initially, when t ∗ < tc∗, the
at one particular point in time and so it does not necessarily reflect a temperature of the nanofluid is in steady-state and so the fv = 0.006 plot
collector’s overall operating efficiency. We define the overall efficiency in Fig. 3 equivalently illustrates the system dynamics. In this region, T ∗
of this solar collector during a specific time interval, as the ratio of the increases linearly as the nanofluid flows through the collector. How-
net amount of energy that exits the system to the overall amount of ever, at tc∗ , the incoming solar intensity decreases instantaneously due
incoming radiation entering the system during that period, i.e., to cloud cover and the nanofluid’s temperature profile enters into a
period of rapid transition where ∂T ∗/ ∂t ∗ ≠ 0 (i.e., when
Eo∗
η= ∗ , tc∗ ⩽ t ∗ ⩽ x ∗/ u ∗ + tc∗) and the nanofluid’s temperature decreases linearly
t
LCA 2πR ∫0 m Gs∗ (t ∗)dt ∗ (24) with time. Even though time-dependence in the source term is in-
stantaneous, its effect on the nanofluid’s temperature at the outlet lasts
where Eo∗
is the overall amount of energy that exits the system during for 34 s after tc∗ . We note that a steady-state model would not capture
the time interval 0 ⩽ t ∗ ⩽ tm∗ , i.e., the nanofluid’s temperature accurately for this period, thus empha-

tm
sising the value of the time-dependent model for real-world applica-
Eo∗ = ∫0 ρnf cp, nf u ∗πR2 (T ∗ (L, t )−TI∗ )dt ∗. (25) tions where the solar intensity is constantly changing. After this tran-
sitional period, (i.e., when t ∗ > x ∗/ u ∗ + tc∗) the temperature profile
We arbitrarily choose tm∗ such that in Scenario 1, tm∗ = 135 s (i.e., 4u ∗/ L ), enters a new linear steady-state regime.
and in Scenario 2 tm∗ = 86400 s (i.e., one day). Also, in Scenario 1 we Fig. 5 shows the temperature versus time of day at five different

Table 3
Dimensionless parameter values for an uncoated solar collector for varying input temperatures for three different particle volume fractions
(fv = 0, fv = 0.0005, fv = 0.006) .

TI∗ = 25 °C TI∗ = 100 °C TI∗ = 200 °C

1/Pe (0.4316, 0.4215, 0.4366) × 10−7 (0.3804, 0.3716, 0.3848) × 10−7 (0.3204, 0.3129, 0.3241) × 10−7
γ (0.0738, 0.0278, 0.0198) (0.0738, 0.0278, 0.0198) (0.0738, 0.0278, 0.0198)
φ (0.0108, 0.0225, 0.0289) (0.0109, 0.0227, 0.0293) (0.0103, 0.0215, 0.0277)
τ4 (0.0789, 0.0297, 0.0212) (0.1747, 0.0658, 0.0468) (0.4517, 0.1701, 0.1210)

77
G.J. O’Keeffe et al. Solar Energy 174 (2018) 73–82

steady-state, however, it spikes at t ∗ = tc∗ reaching values greater than 1.


This non-physical result (η > 1) occurs because the instantaneous effi-
ciency is ill-defined for time-dependent incoming solar intensity. We
remind the reader that Duffie and Beckman (2013) define ηt (t ∗) as the
ratio of usable thermal energy to incident solar energy. At t ∗ = tc∗, the
incident solar energy decreases instantaneously, however the nanofluid
at the collector’s outlet was heated by the previous elevated solar in-
tensity value. After t ∗ = tc∗, the efficiency values decrease before
reaching a new steady-state when all of the nanofluid that had been
heated by the larger solar intensity has flowed out of the system. The
initial steady-state efficiency (when Gs∗ (t ∗) = Gm∗ ) is larger than the new
steady-state efficiency (when Gs∗ (t ∗) = Gm∗ /2 ) across all inlet tempera-
tures—the collector is less efficient when the intensity of incoming solar
radiation is reduced. Although we note in both scenarios that the
NDAPSC coated by the heat-mirror emits much less black-body radia-
tion and hence its steady-state instantaneous efficiency is not very
sensitive to the changes in Gs∗ (t ∗) . In Scenario 2 we note that the un-
coated collectors have negative instantaneous efficiency scores when
the intensity of the incoming solar radiation is too low and
τ 4 > γ + Gs (t ) . Of course, these collectors would be switched off under
Fig. 4. Temperature at x ∗ = 0.25L, x ∗ = 0.5L, x ∗ = 0.75L and x ∗ = L versus time
for Scenario 1 where fv = 0.006 and TI = 200 ° C. such operating conditions.
Fig. 7 shows the overall NDAPSC efficiency versus particle volume
fraction when TI∗ = 200 °C (solid lines), TI∗ = 150 °C (dotted lines),
TI∗ = 100 °C (dot dashed lines), and a heat-mirror coated NDAPSC when
TI∗ = 150 °C (dashed lines) for (a) Scenario 1 and (b) Scenario 2. Effi-
ciency increases rapidly as nanoparticles are initially added to the base-
fluid before plateauing as Π → 1. This is in agreement with previous
theoretical and experimental studies of nanofluid-based direct absorp-
tion solar collectors (Tyagi et al., 2009; Taylor et al., 2011; Khullar
et al., 2012; Xu et al., 2015; Li et al., 2016; O’Keeffe et al., 2016,
2018a,b). In Scenario 1 the uncoated NDAPSC outperforms the heat-
mirror coated NDAPSC. Meanwhile, in Scenario 2, the heat-mirror
coated collector is the most efficient across all particle volume fractions.
The difference between these two results highlights incoming solar
radiation’s affect on collector performance. The mean solar intensity,
Gs∗ , differed significantly across both scenarios: in Scenario 1,
Gs∗ = 750 W m−2 , while in Scenario 2, Gs∗ = 304.6 W m−2 . Since the heat-
mirror is more thermally efficient (i.e., it emits less black-body radia-
tion) than the glass envelop, when incoming solar radiation is lower
and thermal emissions are proportionally larger, the heat-mirror
coating is a superior design choice. However, the glass envelop is more
optically efficient (i.e., it transmits more solar radiation), so when Gs∗ is
higher, this increased optical efficiency offsets thermal losses.
Fig. 5. Temperature at x ∗ = 0, x ∗ = 0.25L, x ∗ = 0.5L, x ∗ = 0.75L and x ∗ = L Fig. 8 shows the overall collector efficiency versus inlet temperature
versus time of the day in Scenario 2, for fv = 0.006 and TI = 200 °C . of an uncoated NDAPSC (solid lines), and a heat-mirror coated NDAPSC
(dashed lines) for (a) Scenario 1 and (b) Scenario 2. As expected, the
positions along the collector for Scenario 2, when fv = 0.006 and solar collector efficiency decreases with increasing inlet temperature;
TI = 200 °C . The nanofluid’s temperature falls/rises approximately lin- although, the efficiency of the uncoated collector decreases more ra-
early as it flows through the receiver, however, the gradient of this pidly than the coated collector. At lower operating temperatures (when
temperature decrease/increase varies throughout the day. As expected, optical efficiency is desirable) the uncoated NDAPSC is more efficient
the solar collector’s outlet temperature, T ∗ (x ∗ = L) , increases with in- than the heat-mirror coated NDAPSC, and at higher operating tem-
creasing radiative intensity and decreases with decreasing radiative peratures (when thermal efficiency is desirable) the coated NDAPSC is
intensity. From 06:40 pm to 05:20 am the solar collector loses heat more efficient—this result holds across both scenarios. In Scenario 1,
because incoming radiation is too small to overcome thermal losses, the efficiency of both collectors is equivalent at TI∗ = 120 °C , while in
i.e., Gs (t ) + γ < τ 4 . This inequality (which comes from (22)), is ex- Scenario 2 these efficiencies are equal at TI∗ = 68.5 °C. Efficiency values
tremely useful as it allows an NDAPSC’s daily operation cycle to be are lower and the difference between the uncoated and coated collec-
informed by weather forecasts. When Gs (t ) + γ > τ 4 (i.e., from tors is more pronounced in Scenario 2, which is due to the radiative
05:20 am to 06:40 pm), incoming radiation overcomes thermal losses intensity being lower, on average. If the collector in Scenario two was
and the nanofluid heats up as it flows through the receiver; an NDAPSC only operational when ηt (t ∗) > 0 its overall efficiency would be larger.
should only be operational during such circumstances. Fig. 9 shows daily energy output of the solar collector versus re-
Fig. 6 shows the instantaneous efficiency ηt (t ∗) versus time for ceiver radius for an uncoated collector (solid line), and a heat-mirror
TI∗ = 200 °C (solid lines), TI∗ = 150 °C (dotted lines), TI∗ = 100 °C (dot coated collector (dashed line) when (a) TI∗ = 100 °C and (b)
dashed lines), and a heat-mirror coated NDAPCS when TI∗ = 150 °C TI∗ = 200 °C. The volume flow rate is kept constant at 3.42 × 10−4m3s−1,
(dashed lines) for (a) Scenario 1 and (b) Scenario 2. This figure de- and so u ∗ decreases with R2 . Also, the aperture width is fixed, so CA
monstrates why an instantaneous measure of efficiency may be mis- decreases linearly with R. Therefore, the dimensionless quantities γ and
leading. In Scenario 1, the instantaneous efficiency is initially at a τ 4 increase linearly with R which implies that as R increases, radiative

78
G.J. O’Keeffe et al. Solar Energy 174 (2018) 73–82

Fig. 6. Instantaneous efficiency versus time for TI∗ = 200 °C (solid lines), TI∗ = 150 °C (dotted lines), TI∗ = 100 °C (dot-dashed lines), and a heat-mirror coated NDAPCS
when TI∗ = 150 °C (dashed lines) for (a) Scenario 1 and (b) Scenario 2.

Fig. 7. Overall efficiency versus particle volume fraction when TI∗ = 200 °C (solid lines), TI∗ = 150 °C (dotted lines), TI∗ = 100 °C (dot-dashed lines), and a heat-mirror
coated NDAPSC when TI∗ = 150 °C (dashed lines) for (a) Scenario 1 and (b) Scenario 2.

Fig. 8. Overall efficiency versus inlet temperature for an uncoated NDAPSC (solid lines), and a heat-mirror coated NDAPSC (dashed lines) where fv = 0.006 . (a)
Scenario 1 and (b) Scenario 2.
79
G.J. O’Keeffe et al. Solar Energy 174 (2018) 73–82

Fig. 9. Daily energy output of the solar collector versus receiver radius for an uncoated collector (solid line), and a heat-mirror coated collector (dashed line) when
(a) TI∗ = 100 °C and (b) TI∗ = 200 °C . The volume flow rate is kept at a constant value of 3.42 × 10−4m3s−1, and the aperture width is also kept constant.

Fig. 10. Daily energy output of the solar collector versus nanofluid particle volume fraction for uncoated collector (solid line), and heat-mirror coated collector
(dashed line) when (a) TI∗ = 100 °C and (b) TI∗ = 200 °C .

heat losses become more significant. Physically, this is due to the re- similar to the efficiency results reported in Fig. 7: as nanoparticles are
ceiver’s suface area (the boundary where thermal emissions occur) added to the base-fluid, we observe a large increase in Eo∗; however,
being directly proportional to R. As expected, Fig. 9 demonstrates that these initial performance enhancements plateau as the nanoparticle
as R increases, collector efficiency decreases. The uncoated collector is concentration continues to increase. As Π → 1, all of the available in-
more efficient than the coated collector (except when TI∗ = 100 °C and coming radiation has already been absorbed and so additional nano-
R < 0.018 m ), and the uncoated collector’s efficiency is much more particles do not improve collector performance. Daily energy output is
sensitive to changes in R since its surface has a higher emissivity. To put larger in Fig. 10(a) when TI∗ = 100 °C , than in Fig. 10(b) when
these figures in perspective, it is estimated that a US household uses on TI∗ = 200 °C: since thermal losses are larger at higher operating tem-
average 10,932 kWh of electricity every year (Energy, 2015), this is peratures, Eo∗ decreases as TI∗ increases. However, the coated collector’s
approximately equivalent to 1.08 × 108 Joules per day. Figs. 9 and 10 daily energy output is less sensitive to changes in TI∗ than the uncoated
show that the optimum daily energy output from Scenario 2 is roughly collector’s daily energy output since the coated collector is more ther-
equivalent to the daily energy consumption of eight households. mally efficient.
Fig. 10 shows the daily energy output of the solar collector versus
nanofluid particle volume fraction for an uncoated collector (solid line),
and a heat-mirror coated collector (dashed line) when (a) TI∗ = 100 °C 4. Conclusions
and (b) TI∗ = 200 °C. This figure shows results which are qualitatively
This paper proposed a time-dependent, three-dimensional model for

80
G.J. O’Keeffe et al. Solar Energy 174 (2018) 73–82

Fig. 11. Error arising from the assumption that the pipe is running at a steady-state where fv =0.006, R = 0.035m, L = 8m, and CA = 71.43.

the efficiency of an NDAPSC under a turbulent flow regime. The model parameters to assess their effect on collector performance. The NDAPSC
consisted of a system of equations: a partial differential equation de- model showed that the overall energy output decreased as R increased,
scribing the conservation of energy, and a time-dependent radiative and furthermore, the parameter space exploration showed qualitative
transport equation describing the propagation of solar radiation agreement with existing NDAPSC models (Khullar et al., 2014; Li et al.,
through the nanofluid. Writing the model in dimensionless form re- 2016; Menbari et al., 2016; O’Keeffe et al., 2018b): efficiency increased
vealed four controlling dimensionless numbers: one describing the re- with increasing nanoparticle volume fraction, and decreased with de-
lative importance of conduction and advection and three describing the creasing flow rates or increasing inlet temperature; heat-mirrors
heat loss to the surroundings. Realistic parameter values were applied sometimes (but not always) enhance collector efficiency.
to reduce the model further and this indicated that two of the di- Scenario 1 highlighted the superiority of a time-dependent model
mensionless groups had a much lesser impact on the performance of the over a steady state model. The system’s temperature entered a period of
solar collector. In Section 2.6 we obtained an analytical expression for rapid transition immediately after a cloud passed over the re-
the temperature in the collector by solving the dimensionless con- ceiver—none of the existing steady-state models can accurately predict
servation of energy equation via the method of characteristics. This an NDAPSC’s temperature profile while the solar intensity is rapidly
expression for the temperature was then used to obtain collector effi- changing, thus emphasising the value of the model proposed in this
ciency and assess the collector performance under various operating paper.
conditions in Section 3.
Although the model is presented in a generalised time-dependent Acknowledgements
form, for demonstration purposes we presented two realistic time-de-
pendent scenarios. Scenario 1 demonstrated dynamic cloud cover, and G. J. O’Keeffe acknowledges the support of the Irish Research
Scenario 2 demonstrated the variation of solar intensity at different Council (GOIPG/2014/887), and the Mathematics for industry network
times of the day. We used Scenario 1 to highlight how an instantaneous (ECOST/STSM/TD1409/290216/071429). S. L. Mitchell and V. Cregan
measure of efficiency may be misleading and lead to non-physical re- acknowledge the support of the Mathematics Applications Consortium
sults (η > 1), and we used Scenario 2 to illustrate how weather fore- for Science and Industry funded by the Science Foundation Ireland (12/
casting can be used to decide when to begin and end an NDAPSCs daily IA/1683). T. G. Myers acknowledges the support of a Ministerio de
operation cycle. We also varied several of the system’s physical Ciencia e Innovaciòn (MTM2014-56218).

Appendix A

In Scenario 2, since the nanofluid flows through the collector at a velocity of 0.237 m s−1, it is only in the receiver for 33.76 s, which is a
relatively short duration compared to the overall length of a day. The solar intensity given by (21) changes on a time-scale of hours rather than
seconds, i.e., it is approximately constant over any period of 33.76 s. Thus, we assume that the system is approximately steady-state, i.e.,
∂T
≈ 0.
∂t (26)
Using this assumption conservation of energy in the system given by (17) reduces to
Tx = Gs (t ) + γ −τ 4. (27)
Integrating both sides of (27) and applying the initial condition, T (x = 0) = 0 , yields
T (x , t ) = (Gs (t ) + γ −τ 4 ) x . (28)
The error associated with using this expression of temperature rather than the full expression, (21), is given by

81
G.J. O’Keeffe et al. Solar Energy 174 (2018) 73–82

T −T ⎞
Error(%) = 100 ⎛ 1 ⎜, ⎟

⎝ T1 ⎠ (29)
where T1 is the temperature given by (21), and T is the approximation given by (28). Fig. 11 shows that (28) works well over the course of a
day—max(|Error|) < 0.05%.

References Therm. Eng. 104, 176–183.


O’Keeffe, G.J., Mitchell, S.L., Myers, T.G., Cregan, V., 2016. The effect of depth-dependent
velocity on the performance of a nanofluid-based direct absorption solar collector. In:
Cregan, V., Myers, T.G., 2015. ‘Modelling the efficiency of a nanofluid direct absorption European Consortium for Mathematics in Industry’. Springer, pp. 327–334.
solar collector’. Int. J. Heat Mass Transf. 90, 505–514. O’Keeffe, G.J., Mitchell, S.L., Myers, T.G., Cregan, V., 2018a. Modelling the efficiency of a
Duffie, J.A., Beckman, W.A., 2013. Solar Engineering of Thermal Processes. John Wiley & low-profile nanofluid-based direct absorption parabolic trough solar collector. Int. J.
Sons. Heat Mass Transf. 126, 613–624. <http://www.sciencedirect.com/science/article/
Ellabban, O., Abu-Rub, H., Blaabjerg, F., 2014. Renewable energy resources: current pii/S0017931018307221> .
status, future prospects and their enabling technology. Renew. Sustain. Energy Rev. O’Keeffe, G.J., Mitchell, S.L., Myers, T.G., Cregan, V., 2018b. Modelling the efficiency of a
39, 748–764.. <http://www.sciencedirect.com/science/article/pii/ nanofluid-based direct absorption parabolic trough solar collector. Sol. Energy 159,
S1364032114005656> . 44–54. <http://www.sciencedirect.com/science/article/pii/
Elperin, T., Kleeorin, N., Rogachevskii, I., 1996. ‘Turbulent thermal diffusion of small S0038092X17309465> .
inertial particles’. Phys. Rev. Lett. 76 (2), 224. Otanicar, T.P., Phelan, P.E., Golden, J.S., 2009. ‘Optical properties of liquids for direct
Energy Information Administration, 2015. How much electricity does an American home absorption solar thermal energy systems’. Sol. Energy 83 (7), 969–977.
use?. <https://www.eia.gov/tools/faqs/faq.cfm?id=97&t=3>. Rakić, A.D., 1995. ‘Algorithm for the determination of intrinsic optical constants of metal
Fan, J.C., Bachner, F.J., 1976. ‘Transparent heat mirrors for solar-energy applications’. films: application to aluminum’. Appl. Opt. 34 (22), 4755–4767.
Appl. Opt. 15 (4), 1012–1017. Sawin, J.L., Seyboth, K., Sonntag-O’Brien, Sverrisson, F. et al., n.d. Renewables 2017
Giovannetti, F., Föste, S., Ehrmann, N., Rockendorf, G., 2014. ‘High transmittance, low global status report. <http://www.ren21.net/wp-content/uploads/2017/06/17-
emissivity glass covers for flat plate collectors: applications and performance’. Sol. 8399_GSR_2017_Full_Report_0621_Opt.pdf>.
Energy 104, 52–59. Solutia, n.d. Therminol®VP-1’, URL <http://twt.mpei.ac.ru/tthb/hedh/htf-vp1.pdf>
Khullar, V., Tyagi, H., Hordy, N., Otanicar, T.P., Hewakuruppu, Y., Modi, P., Taylor, R.A., (accessed: 28/02/2017).
2014. ‘Harvesting solar thermal energy through nanofluid-based volumetric absorp- Taylor, R.A., Phelan, P.E., Otanicar, T.P., Adrian, R., Prasher, R., 2011. ‘Nanofluid optical
tion systems’. Int. J. Heat Mass Transf. 77, 377–384. property characterization: towards efficient direct absorption solar collectors’.
Khullar, V., Tyagi, H., Phelan, P.E., Otanicar, T.P., Singh, H., Taylor, R.A., 2012. ‘Solar Nanoscale Res. Lett. 6 (1), 1–11.
energy harvesting using nanofluids-based concentrating solar collector’. J. Tyagi, H., Phelan, P., Prasher, R., 2009. ‘Predicted efficiency of a low-temperature na-
Nanotechnol. Eng. Med. 3 (3), 031003. nofluid-based direct absorption solar collector’. J. Sol. Energy Eng. 131 (4), 041004.
Kimball, H.H., 1935. ‘Intensity of solar radiation at the surface of the earth, and its University of Oregon Solar Radiation Monitoring Laboratory, 2015. Archival solar data
variations with latitude, altitude, season, and time of day’. Mon. Weather Rev. 63 from Seattle, Wahington, June. <http://solardat.uoregon.edu/SelectArchival.
(1), 1–4. html>.
Kolb, G.J., 2011. An evaluation of possible next-generation high-temperature molten-salt Veeraragavan, A., Lenert, A., Yilbas, B., Al-Dini, S., Wang, E.N., 2012. ‘Analytical model
power towers. Sandia National Laboratories, Albuquerque, NM, Report No. for the design of volumetric solar flow receivers’. Int. J. Heat Mass Transf. 55 (4),
SAND2011-9320. 556–564.
Li, Q., Zheng, C., Mesgari, S., Hewkuruppu, Y.L., Hjerrild, N., Crisostomo, F., Xie, W.T., Dai, Y.J., Wang, R.Z., 2011. ‘Numerical and experimental analysis of a point
Rosengarten, G., Scott, J.A., Taylor, R.A., 2016. ‘Experimental and numerical in- focus solar collector using high concentration imaging pmma fresnel lens’. Energy
vestigation of volumetric versus surface solar absorbers for a concentrated solar Convers. Manage. 52 (6), 2417–2426.
thermal collector’. Sol. Energy 136, 349–364. Xu, G., Chen, W., Deng, S., Zhang, X., Zhao, S., 2015. ‘Performance evaluation of a na-
Martinopoulos, G., Missirlis, D., Tsilingiridis, G., Yakinthos, K., Kyriakis, N., 2010. ‘Cfd nofluid-based direct absorption solar collector with parabolic trough concentrator’.
modeling of a polymer solar collector’. Renewable Energy 35 (7), 1499–1508. Nanomaterials 5 (4), 2131–2147.
Menbari, A., Alemrajabi, A.A., Rezaei, A., 2016. ‘Heat transfer analysis and the effect of
CuO/water nanofluid on direct absorption concentrating solar collector’. Appl.

82

You might also like