Sie sind auf Seite 1von 6

Performance figures for the analysis of heat rejection

systems

Bjrn Nienborg
1
, Hannes Fugmann
1
, Lena Schnabel
1
, Peter Schossig
1
, Sebastian
Wittig
2
, Gregor Trommler
2
, Antoine Dalibard
3
1
Fraunhofer ISE (Institute for Solar Energy Systems), Heidenhofstr. 2,79110 Freiburg,
Germany
Email: bjoern.nienborg@ise.fraunhofer.de
Phone: +49 - 761 - 4588 5883
2
ILK Dresden www.ilkdresden.de
3
HfT Stuttgart www.hft-stuttgart.de

Introduction
Past research projects have shown that there are a few typical issues which lead to
deficient operation of solar thermal cooling systems (STCS). Within the project
SolaRck(www.solarueck.de) 3 sorption machine manufacturers and 3 research
institutes cooperate to tackle two of them: the inaccurate design of the heat rejection
subsystem and the suboptimal operation strategy of the overall system. For the first
point (only dry coolers are considered here) measurement data are analyzed with the
goal to derive optimization approaches for the design and a rating procedure for
cooling tower performance. Concerning the second issue optimized generic operation
strategies shall be developed with support of system simulations.
This paper focusses on the presentation of the fundamental work for the first point: the
review and definition of suitable evaluation figures for dry coolers.

Comparing the performance of dry coolers
For the comparison of various types of dry coolers and the evaluation of measurement
data a number of different operation conditions have to be taken into account. These
conditions (e.g. mass flows, temperatures) change over time and most likely vary
among heat rejection systems. A comparison method is only valuable, if these
differences are considered. In the following, data from 4 sources were used:
Data calculated with the software CoilDesigner (J iang, Aute et al. 2006) for a
small dry cooler; air and water mass flow as well as both inlet temperatures are
varied (DC1)
2 sets of monitoring data from the German project SolCoolSys
(www.solcoolsys.de);air mass flow and both inlet temperatures are varied (DC2
and DC3); since the dry cooler is operated with a glycol mixture in these

installations, the heat capacity is measured in the water circuit after a separating
heat exchanger.
Lab data measured by German manufacturer Thermofin (www.thermofin.de); air
and water mass flow as well as both inlet temperatures are varied (DC4)
Based on this data and analytical simplifications, we developed one method (extended
effectiveness-NTU method), which fulfills these requirements and is therefore a strong
tool in comparing the performance of different heat rejection systems.
To evolve a better understanding of the difficulties arising when comparing the
performance, we first present three methods, which cant depict the whole complexity,
but use existing performance figures.
The first method describes the specific electricity consumption versus the cooling ratio.
The specific electricity consumption is defined as the quotient of the electrical power
consumption of the circulating pump and the fan (

in [W
el
]) and the rejected heat (


in [W
th
]). The cooling ratio

is the quotient of temperature range of water (

in
[K]) and the inlet temperature difference ( =
,

,
in [K]):

/ in [K/K]
(1)
In Figure 1 the used data are plotted. An increase in cooling ratio coincides largely with
an increase in specific electricity consumption. A rating based on this data leads to the
assumption that DC1 is best, followed by DC2 and DC3. It is not possible to classify
DC4.

Figure 1: Comparison of dry cool ers based on method 1
One advantage of this method is the clarity of the used parameters. An effective heat
rejection system has a low specific electricity consumption at high cooling ratios. An
important disadvantage of this method is that for high values of a high value for


cooling ratio is reached at a low value for specific electricity consumption. This
circumstance leads to a better rating of heat rejection systems and operation with high
, although this operational condition should be avoided, as it coincides with a poor
performance of the STCS. In Figure 1 this is the case for the data of DC1 and DC2.
The medium of DC1 is 18.2K, the medium of DC2 is 9.9K. Therefore the better
rating of DC1 in Figure 1 is due to the applied method and not necessarily due to a
better heat exchanger.
The second method can somehow compensate this disadvantage, by using the
cooling potential as one parameter of performance and the specific electricity
consumption as the second parameter. The cooling potential is defined as the quotient
of ambient temperature (
0
in [K]) and arithmetic mean temperature of the water
(
,
in [K]). In an ideal dry cooler
,
is equivalent to
0
. This means on
the one hand that no exergy (useful energy) is passed to the ambient; on the other
hand this can only be reached with an infinite heat transfer surface. Therefore the
cooling potential is always less than 1. The performance of a dry cooler can be rated
by how close the cooling potential is to 1. Figure 2 shows that an increase in cooling
potential coincides largely with an increase in specific electricity consumption. The
scattering of data is due to the fact that a fixed value of cooling potential can be
reached by different operation conditions (changing mass flows, changing inlet
temperatures). This results in different specific electricity consumption for the same
value of cooling potential. However as the operation conditions of DC1 and DC2 are
comparable, a rating between these two dry coolers based on Figure 2 is still possible.

Figure 2: Comparison of dry cool ers based on method 2
Method 3 uses the overall heat transfer coefficient ( in [W/(K m)]) as a measure of
performance. This measure is plotted in Figure3 versus the specific electricity
consumption. An increase in coincides largely with an increase in the specific

electricity consumption. But due to changes in water mass flow it is not possible to
define a direct dependence of specific electricity consumption to . The reason is that
changes in water mass flow result in changes in heat rejected (

). The electrical power


and the overall heat transfer coefficient depend only to a minor degree on the water
mass flow. Therefore for a given value of , a variation of

leads to a change in the


specific electricity consumption. The scattering in Figure3 is much more pronounced in
DC1, than in the other data, as the water mass flow is varied over a wide range for
DC1.

Figure 3: Comparison of dry coolers based on method 3; the low calculated values for DC2 and
3 are a consequence of the separating heat exchanger in the cool ing ci rcuit
A comparison between DC2 and DC3 is possible as the water mass flow is constant.
Once again DC2 is rated better, as the value of is higher for the same specific
electricity consumption.
Evidently some constraints have to be set to the operation conditions in order to
compare different dry coolers according to the presented methods. Hence a more
analytical approach is chosen to keep the number of constraints (resp. the number of
necessary parameters) small.

In Method 4 we develop the effectiveness-NTU method (e.g (Krger 2004)) further. It
results in a method that describes the electrical power of the dry cooler with respect to
solely two parameters. These are the operation number (), defined as =

/(

,
) in [kg/s] and the cooling ratio

.
The heat exchanger effectiveness () is in general a function of number of transfer
units (), the heat capacity rate ratio (

) and the flow arrangement for a direct-


transfer type heat exchanger ((Shah and Sekulic 2003)). The flow arrangement is fixed
for a given dry cooler (DC), so the function can be written as:


= (,

, DC) (2)
The function will be used throughout the paper, always describing a not explicitly
known function. It can change from formula to formula. For some flow arrangements
can be expressed explicitly as a function of and

. For other flow


arrangements can be expressed only implicitly, but for our case this is sufficient.
So is a not explicitly known function.
= (,

, DC). (3)
For dry coolers is approximately a function of

and this results in:


= /

= (

, , DC) (4)
Further and

can be easily expressed as:



= (

, , DC)

= (

, , DC)
(5)
Combining Equation (2) to (5) yields

= (

, , DC) in [kg/s] (6)


In addition the water mass flow is directly given by a function:

= (

, ) in [kg/s] (7)
As the electrical power is determined only by the mass flows, we showed

= (

, , DC) in [W
el
] (8)

This approach enables us to describe the electrical power needed to run pumps and
fans for a given dry cooler by solely knowing the value of and

. Figure 5 shows the


measurement data for constant

= 0.34. For a given amount of heat, that has to be


rejected at given temperature differences and

one can easily determine the


electrical power which is needed to get to these operational conditions. For 1.25
the electrical power consumption of DC1 equals 56.2W ( = 1.26), for DC2 129.3W
( = 1.38), for DC3 450.8W ( = 1.29), and for DC4 1226W ( = 1.17). Thus this plot
can be used for the rating of cooling towers. However it contains no information on size
and investment costs. Normalizing the electrical power and B versus the size or costs
can solve this issue. Since the reference parameter (e.g. heat exchanger area,
volume, weight, investment cost) has a strong impact on rating this topic will be further
analyzed.


Figure 4: Comparison of dry cool ers based on method 4; eta_c= 0.34 = const.

Conclusions
In the course of this paper different approaches for the comparative evaluation of dry
coolers are presented and their advantages and draw-backs are discussed. Based
on the constraints of the first three methods, an extended effectiveness-NTU method
was developed. It appears very promising as it requires only relevant operation data of
the dry coolers. In the future we still need to analyze the possibility of normalizing the
results to the size or investment cost. We also plant to evaluate the applicability of the
developed approach to wet cooling towers.
Concerning the optimized generic operation strategies, the potential of different
optimization approaches will be studied. These may be based on existing
approaches to improved control strategies for STCS e.g. studied by (Albers 2009)
previously. Additionally innovative approaches for operation strategies which are
have found their way into other fields of technology, will be analyzed.


References:
Albers, J . (2009). Solar-driven Adsorption Chiller Controlled by hot and Cooling
Water Temperature. Otti Solar Air Conditioning Conference, Palermo, Italy.
J iang, H., V. Aute, et al. (2006). "CoilDesigner: a general-purpose simulation and
design tool for air-to-refrigerant heat exchangers." International J ournal of
Refrigeration 29(4): 601-610.
Krger, D. G. (2004). Air-cooled Heat Exchangers and Cooling Towers, PennWell
Books.
Shah, R. K. and D. P. Sekulic (2003). Fundamentals of heat exchanger design,
Wiley.

Das könnte Ihnen auch gefallen