Vous êtes sur la page 1sur 43

Chapter 2

Controlled Resistive Heating of Carbon Fiber Composites


Temperature control is the key to eective resistive heating for composite rigidization. The ability to monitor and even control the temperature provides an active control strategy in prescribing and predicting the stiness of the hardened composite. Further, understanding how the material responds to an electrical input determines both the type of control to be used as well as the level of control that can be attained. This chapter investigates how temperature control of carbon-ber reinforced polymer composites is implemented. The material was investigated experimentally and modeled analytically for use in the nal control scheme. A PID feedback controller was applied to the resistive heating process and the control gains were then experimentally tuned. Infrared imaging of the composite during heating was also performed as a method for visualizing the heating process.

2.1

Introduction to Resistive Heating

When Georg S. Ohm (1787-1854) published Die galvanische Kette mathematisch bearbeitet in 1827, he described the theory and applications of electric current. For his achievements, the German scientists name has been forever attributed to electrical science terminology: Ohms Law (2.1) states the proportionality of current and voltage in a resistor and the SI

18

unit of resistance is the Ohm () [32]. V = IR. (2.1)

Ohms work provides a direct correlation between the potential voltage drop, V , across a resistive (or Ohmic ) material, R, and the electric current, I , owing through the medium. In doing so, Ohms Law is a fundamental concept that helped establish the basis of modern electrical theory. James Joule (1818-1889) coupled Ohms Law with his own endeavors in relating heat to mechanical work. Joule worked with Lord Kelvin in developing the absolute temperature scale (Kelvin, K) and was also acknowledged for his contributions that eventually led to the First Law of Thermodynamics. His experiments initiated the concept of the mechanical equivalence of heat, which relates the energy required to raise the temperature of water 1 F. Accordingly, the SI unit of work was named the Joule (J). However, Joule might be best remembered for his discovery of the relationship, appropriately named Joules Law, between current ow and heat dissapation in a resistive element (2.2). , within a resistive Specically, Joule found that the rate of thermal energy generated, E material, = I 2 R, E (2.2)

is proportional to the square of the current, I , and directly proportional to the resistance, R [33, 34]. Together, Georg Ohms and James Joules accomplishments close the gap between electrical energy input (either as a voltage potential or an electric current) and thermal energy output. As a result, resistive heating can provide both undesired and desired heat generation. Because of Joules Law, resistive heating occurs whenever an electric current passes through a resistive material. The design of electronic circuitry must account for this phenomena or the risk of overheating, even melting, the hardware becomes a reality. On the other hand, Joules law sets forth the notion that the amount of thermal energy generated can be controlled by the inherent resistance of the heating element as well as the applied electric current. The selection of the input, i.e. current, voltage, or even resistance, to obtain a desired temperature rise is the most common use for resistive heating. Appliances like electric stove tops, hairdryers, curling irons, and electric blankets employ resistive heating to raise the temperature of heating elements. Internal resistive heating of carbon-ber reinforced materials provides a way to a 19

new method of rigidization. If current passing through the bers can generate sucient heat, then it is thought that the localized temperature increase can be used to cure the adjacent thermosetting resin. As Joule demonstrated, this process is an active one; the temperature of the composite can be controlled via the supplied electrical input. Therefore, the requirement for a given temperature to be obtained during the rigidization process is merely a function of the electrical energy supplied to the material.

2.1.1

The Resistive Nature of CFRP Materials

In order for carbon ber reinforced polymers to be rigidized by resistive heating, the composite must have the appropriate electrical properties. When an electric eld is applied to a material, the motion of electric charges within the material generates current ow. However, all materials exhibit some resistance, R, to this charge motion. Resistance, then, depends on both the resistive nature of a given material, called electrical resistivity, and the geometry of the material l R= . A (2.3)

Equation 2.3 shows that the resistance of a heating element is proportional to its length, l, and inversely proportional to its cross-sectional area, A. Further, the electrical resistivity, , for a given material is a function of temperature. The electrical conductivity, e , is a measure of the materials ability to allow electrical current to ow and is dened as the reciprocal of resistivity [3] e = 1 . (2.4)

Table 2.1: Room Temperature Electrical Conductivities of Common Engineering Materials [3] Metals and Alloys e , ( m)1 Nonmetals e , ( m)1 Silver 6.3 x 107 Graphite 105 (average) 7 Copper, commercial purity 5.8 x 10 Germanium 2.2 Gold 4.2 x 107 Silicon 4.3 x 104 Aluminum, commercial purity 3.4 x 107 Polyethylene (PE) 1014 Polystyrene (PS) 1014 Diamond 1014 Quantied as -m1 , electrical conductivity (and thus resisivity) is an inherent material property that does not depend on geometry. Both conductivity and resistivity are functions are temperature, however. The dierence in conductivity varies greatly between 20

materials and helps categorize materials as conductors, insulators, and semiconductors. Table 2.1 illustrates that pure metals have the highest conductivities, which explains their use as electrical wiring materials. On the opposite end of the spectrum, the electrical insulators such as polyethylene (PE), polystyrene (PS), and diamond have very small conductivity values. Between these extremes, semiconductors such as graphite, germanium, and silicon have conductivities less than metals but much greater than insulators. An in-depth look at the physical properties of both the carbon ber tow and thermosetting resin is discussed in Chapter 3. For now, the primary focus is kept on the method of applying internal resistive heating to resistive materials.

2.1.2

Resistive Heating Experimental Goals

The application of resistive heating to the carbon-ber reinforce polymeric materials of interest required an understanding of how the material heats as well as the electrical power requirements needed to reach curing temperatures. Thermal control was implemented on these materials through a two-step process. First, the material was subjected to open-loop heating tests, in which a constant voltage was applied to the material for a given amount of time. In doing so, techniques for measuring the temperature were established and the power requirements to achieve certain temperatures was recorded. The heating and cooling time constants were also measured as a way of experimentally modeling the heated material. Closed-loop, or feedback, control was then applied to the heating process. In this strategy, a measured temperature was compared to a desired temperature and the input voltage (or current) was regulated in order to reduce the error between the two. Selected for its eectiveness and simplicity, a proportional-integral-derivative (PID) was chosen as the type of control strategy to use. The goals, and resulting success, of this testing were chosen in order to produce an eective, repeatable resistive heating rigidization method for CFRP materials. The prescribed stiening that this method can produce, then, is very much dependent on the ability to control the temperature of the material. The specic detailed objectives and goals for this testing are summarized: Establish temperature measurement of CFRP samples during resistive heating. Measure temperatures achieved for selected input voltages during open-loop testing. Record heating and cooling time constants for use in system identication. 21

Apply feedback control, by way of a PID controller, to the resistive heating process. Tune the controller to achieve accurate and precise temperature control.

2.2

Thermal Control Experimental Setup

To perform temperature observation during resistive heating, the samples were rst xed on each end. Each sample consisted of a 15 20cm length of polymer resin-coated carbon ber tow. An applied voltage (generated in Simulink and dSpace and then amplied by a 20V, 2A HP 6825A Bipolar Power Supply/Amplier) across the sample resulted in current ow between the voltage leads. Omega J-type (Iron-Constantan) thermocouples were placed at three positions along the sample to measure temperature. An additional thermocouple was also used to record the ambient air temperature. Temperatures from the thermocouples were then measured and recorded by an Omega OMK-DAQ-56 data acquisition module and Personal DAQView software. Then, data from this device was transmitted via USB to a personal computer.
OMK-DAQ-56

Simulink

Th1

Th4
i

Th6

V+
dSpace Controller

V-

Thamb

Data

V(t)
Power Supply/Amp Personal Computer

Figure 2.1: Test setup schematic for open-loop resistive heating with temperature monitoring. The temperature was measured at multiple positions in order to detect possible temperature dierences, or gradients, along the length of the material. Since the tow consisted of such ne bers, the placing of the thermocouples (much larger in contrast) required multiple iterations. At rst, untwisted ber tow was position into the xture (2.3). However, the bers were easily separated, allowing the thermocouple to pass through the thickness of the material. Poorly placed thermocouples not only lead to imprecise 22

Th1

Th2

Th3

V+

Composite Sample

V-

Figure 2.2: Resistive heating with temperature monitoring experimental setup. temperature measurements, but in the presence of a feedback control algorithm can cause incorrect amounts of energy to be supplied, resulting in an inaccurate heating method. One solution to this problem was twisting the ber tow samples. This technique kept the bers in a tighter knit conguration and in turn, caused the sample to have a more consistent cross-sectional area.

Figure 2.3: Thermocouple positioning without twisting (left) and with twisting (right). Notice how the tip of the thermocouple protrudes through the tow in the rst picture. Twisting the samples helped to hold the thermocouple in a position to better measured internal temperature. This experimental setup was also designed to measured the amount of current owing through the material. Knowing how much current ows allows for the electrical power to be calculated as well as the resistance of the material to be monitored during heating. To perform these tasks, a current-measuring circuit was designed and built (Figure 2.4). Modeling the sample as a resistor with unknown resistance, Rstrands , the output voltage, Vout , was measured directly during the tests. Knowing the voltage drop across a known

23

R2

R1 Vi Rstrands + Vin + V R3 V+

+
OP177

+ Vout -

Figure 2.4: Current sensing circuit used to measure current ow and material resistance. resistance, R3 , allowed the current owing through R3 (and also the sample, Rstrands ) to be determined. Likewise, the voltage drop, V , across the strands was measured and from that, the resistance of the sample was obtained. Figure 2.5 illustrates a ow-chart type representation of how each unknown value was computed.
Vin Vout R2/R1 R3

Vout =

2 Vin ( R R1 + 1)

(1 + )
Rst R3

Vin Rst = R3 V out

R2 R1

+1 R3

Rst V+ V i

V + =V =

Vout (1 + RR12 )

V = Vin V + i= V+ R3

Figure 2.5: Post-processing circuit and associated equations for indirect current measurement. This testing apparatus (Figure 2.1) was then used to perform open-loop resistive heating on samples of the CFRP tow. However, when closed-loop control was performed the experimental setup was changed to accomodate temperature as a feedback control signal. Specically, four signal conditioners (Omega #CCT-22-0/400C) with a range of 0 400o C were used to provide the cold-junction reference point for the thermocouples and then output 0 10V voltage signals proportional to the measured temperatures. These signal conditioners replaced the OMK-DAQ-56 temperature data acquisition module, which could not output temperature signals. When incorporated into the previous experimental

24

Thermocouple Signal Conditioners

Figure 2.6: Signal conditioners used to measure and output temperature signals required for feedback control. setup (Figure 2.1), the new experimental setup sends measured temperatures back into Simulink/dSpace, where the control algorithm could process them.
*Temperature Feedback

LowPass RC Filtering *PID Controller Algorithm Simulink (Lab PC)

SC1

SC2

SC3

SCamb

IN

Th1

Th2
i

Th3

Thamb

dSpace Controller
OUT

V+

V-

*Control Signal

Power Amplifier

Figure 2.7: Experimental schematic used for feedback temperature control. In order to visualize the procedure of applying a PID control to this system, the Simulink Block diagrams used to control temperature are shown Figures 2.8. In these diagrams, the conversion of measured temperature to control voltage is traced. Using dSpace to accept thermocouple voltages, the input signals were then converted to temperature and the average temperature was computed. The error between the measured temperature and the temperature set point was then scaled by the PID control gains, which generated a control eort power value. Taking into account changes in the ambient temperature (an external disturbance) and using a constant initial resistance value for the 25

Channel 1 Voltage

V1

T1 T2

Ro 1 10
Ro Tset V Tav g Tamb V(t)

Channel 2 Voltage

V2 T3

M Tset Data

Channel 3 Voltage

V3 Tav g

Channel 4 Voltage

V4

Tamb

Channel Outputs from D-Space

Channel 5 Voltage

Temperature Conversion

PID Control

Channel Inputs to D-Space

Current Voltage

Current

Current Monitoring

2 Tset 3 Tavg 4 Tamb


error PID Control

f(t)=(V^2)/R+hAsT_inf sqrt PID Loop 0.083 hAs 1 V

1 Ro

0.4 Kp 1 error Ki 0 Kd du/dt 0.04 1 s 1 PID Control

Figure 2.8: a. Simulink model created to house the temperature control algorithm (top). b. Temperature feedback control loop used to compute the corrective control voltage (middle). c. PID controller (bottom) located within the temperature feedback loop (b). sample, a corrective control voltage was calculated. This signal was then sent out of dSpace, through the power supply/amplier, and into the material.

2.3

Open-Loop Heating Results and Discussion

The resistive heating process was initially examined by inputting a single voltage pulse and recording the temperature response along the strands. The pulse length was selected to allow the material to reach a nal, or steady-state, value. Then the voltage was turned o and the temperature was observed as the material cooled down. Figure 2.9 shows the temperature responses measured at the three locations along the samples for an input voltage of 10V with 3 minutes of heating and 3 minutes of cooling. The measured responses illustrate

26

t2sample52.mat 150 Th1 Th4 Th6

Temperature - C

100

50

50

100

150

200

250

300

350

400

15 10 Vin Vout

Voltage

5 0 -5 -10 0 50 100 150 200 Time - s 250 300 350 400 Max Power = 16.458 W Max Current = 1.6172 A

Figure 2.9: Typical temperature and voltage responses measured for an input voltage of 10V. some important aspects of the heating and cooling properties of this type of composite material. First and foremost, the material experiences an exponential temperature rise (when voltage is turned ON) from an initial temperature, Ti , up to a higher nal temperature, Tf . Cooling is denoted by an exponential decay (when voltage is OFF) from an initially high temperature, Ti , to cooler nal temperature, Tf . Each of these responses are approximated by the following equation T (t) = Aect + B where, A = Ti T f B = Ti (2.6) (2.7) (2.5)

and c represents the exponential growth (heating) or decay (cooling) rate. The corresponding heating and cooling time constants, 1 = , c (2.8)

which refer to how quickly the material responds to a given input, can also be measured from these plots. Values of the heating constants are then later used to rene the predicted response model developed for controlled, open-loop heating. The voltage plot in Figure 2.9 gives additional insight into the power requirements of the sample. Shown as a blue, dotted line, the input voltage, Vin , represents the voltage 27

applied across the strands. In addition, a red, dotted line plots the values of an output voltage, Vout , measured from the post-processing circuit during the experiment. Recall that two additional goals of this test were to measure the current owing through the material and, from that measurement, track changes in the electrical resistance of the material. Returning to Figure 2.9, it is noticed that Vout remains at during the heating process. Examining the top equation in Figure 2.5 further claries how the resistance of the strands changes for a given change in the output voltage, Rst = R3 Vin Vout R2 +1 R1 R3 Rst = G R3 , Vout (2.9)

where G is a constant if all other resistances and the input voltage remain xed. Taking the partial derivative of Equation 2.9 with respect to the measured output voltage yields the following G Rst G = 2 Rst = 2 Vout . Vout Vout Vout (2.10)

As derived, a change in resistance of the sample, Rst , is proportional to a change in the output voltage, Vout . Since Figure 2.9 shows a relatively unchanging measurement of Vout over the length of the heating cycle, we can infer that the resistance of the strands also remains steady. However, a steady resistance is also dependent on the material having a resistivity that does not vary with temperature. The amount of resistance change per temperature increase, then, will also be considered during this test. As it will be shown later, a xed value of resistance greatly simplies the both the simulation of the temperature response as well as the eventual temperature control of the process. Previously mentioned in the goals for this experiment, it was desired to record the maximum temperature (an average of the three thermocouple measurements along the sample) achieved for varying amounts of applied voltage. For each sample, the input voltage was varied (from 0.5V to 10V in 0.5V increments) and the corresponding temperature response was measured (Figure 2.10). It was observed that the temperature of the tow sample was proportional to the square of the applied voltage. Further, by measuring the current with the sensing circuit shown in Figure 2.4, the amount of current to obtain these temperatures was also examined. Again, the temperature was proportional to the square of the current. A second-order polynomial trend-line veries this relationship graphically. The quadratic relationships between temperature and voltage (and current) shown 28

Plot of all Thavg Responses 110 Run1 Run2 Run3 Run4 Run5 Run6 Run7 Run8 Run9 Run10 Run11 Run12 Run13 Run14 Run15 Run16 Run17 Run18 Run19 Run20

100

90

80
Temperature - C

70

60

50

40

30

20

50

100

150

200 Time - s

250

300

350

400

Figure 2.10: Average temperature response of the sample for various input voltages (Run1=0.5V and Run20=10.0V). in Figure 2.11 are consistent with Equation 2.2, which states that the heat produced by an electrical signal in a resistive element is proportional to the square of the current. Further, if Equation 2.1 is substituted into Equation 2.2, the heat generated is equally proportional to the square of the applied voltage = E V R
2

R=

V2 . R

(2.11)

As it will be further shown in later sections, the temperature of the material is directly proportional to the amount of heat generated within it. In addition, the heat generated within the material is proportional to the power supplied to the material. This relation, then, states that the temperature attained is also linearly proportional to the power supplied (Figure 2.12) P = V I = I 2R = V2 . R (2.12)

Heating and cooling time constants for each sample test were also computed from the average temperature response and were later used to help model the heating process. This knowledge of the systems response to a known input was also valuable when applying 29

120 100

120 100 y = 20.508x + 23.985x + 17.329 2 R = 0.9972


2

Temperature - C

Temperature - C
8.0 10.0 12.0

y = 0.6253x + 2.1939x + 18.666 80 60 40 20 0 0.0 2.0 4.0 6.0 Voltage - V R = 0.9968


2

80 60 40 20 0 0 0.5 1 Current - A 1.5 2

Figure 2.11: Measured average maximum temperature per input voltage (a) and current (b) applied.
120 100

Temperature - C

80 y = 5.3283x + 23.903 60 40 20 0 0 5 10 Power - W 15 20 R = 0.9908


2

Figure 2.12: Material temperature exhibits a linear relation to electrical power. feedback control to the heating procedure. From this round of tests, an average heating time constant, h and an average cooling time constant, c , were extracted from the measured data. These values represent the time for the response to reach 1 e1 (63.2%) of its steadystate value [35]. The time constants measured demonstrate no signicant dependence on the applied voltage (and thus, the temperature attained) and were roughly the same for heating (19.2 1.8 seconds) and cooling (20.8 2.6 seconds). A discussion of how these time constants are modeled is later addressed and shows that these values are primarily dependent on the material itself. The resistance of the sample was also mapped versus the temperature reached for each level of applied voltage. It was desired to observe if this property changed greatly or if it could be assumed to remain constant during the heating process. A constant heating process translates into a time-invariant system model for use in feedback control. Otherwise, the control strategy must continually update the resistance value as the temperature

30

30.0 25.0
Time Constants - s

20.0 15.0 10.0 5.0 0.0 0.0 2.0 4.0 6.0 Input Voltage - V 8.0 10.0 12.0 Heating Cooling

Figure 2.13: Measured heating and cooling time constants for open-loop resistive heating tests. changed. Through the circuit described previously, the resistance of the strands were measured indirectly and tracked relative to an initial resistance value (Figure 2.14). Resistance
16 14
Resistance - W

0.6 0.4 Ro = 10 0.2 0 -0.2 -0.4 -0.6 120


Resistance -

12 10 8 6 4 0 20 40 60 Temperature - C

80

100

Figure 2.14: Sample resistance as a function of temperature. does decrease with increasing temperature, but only slightly. At 100o C, the sample resistance has only decreased by 5.25% to 9.48. The change in resistance might be more of a factor at higher temperatures, though in the temperature range presented, it remained stable. While the temperature of the material has yet to be controlled, resistive heating with temperature monitoring was established and a method of positioning the thermocouples was achieved. Further, the steady-state temperature attained was measured as a function of the voltage, current, and power supplied to the sample. It was also observed that the resistance of the material did not change signicantly during these heating tests. Short of feedback control, curing temperatures for UnyteSet resin (160o C-170o C) have yet to be achieved even in open-loop resistive heating. In order for these higher temperatures to be generated, 31

greater amounts of power must be supplied. The resulting experimental setup takes into account the need for increased power supply and in doing so uncovers a better way to measure current ow.

2.3.1

High-Temperature Open-loop Resistive Heating

Expanding the results from the low temperature experiments, it was desired to track the nal temperatures attained for larger input voltages. In thought, the quadratic trend-line shown in Figure 2.11 would be stretched to higher voltages and temperatures. This step required a change of hardware to safely accommodate higher voltages (and currents). Using a Xantrex XHR 300V-3.5A DC Power Supply, the output voltage was triggered via remote signaling. Simulink and dSpace were still used to generate the signal shape and the new power supply then scaled the signal to desired voltage levels. Another benet of using this hardware was that both the output voltage (equivalent to the voltage potential across the sample) and current from the power supply can be directly monitored and recorded as scaled voltage signals. Simply, by using this power supply, higher power can be sourced to the composite sample and the post-processing circuit was eliminated from the picture. A new round of temperature response testing on samples of G40-800 carbon ber coated with Unyte201 polymer resin was conducted for voltages varying from 1 to 18V.
400 350
Temperature - C

400

Temperature - C

300 250 200 150 100 50 0 0

y = 0.9898x + 0.4696x + 23.232 2 R = 0.9985

350 y = 6.2951x + 30.64 300 250 200 150 100 50 0 R = 0.9945


2

10 Voltage - V

15

20

10

20

30 Power - W

40

50

60

Figure 2.15: Power requirements measured during high-temperature, open-loop resistive heating tests. Again, the voltage and power were monitored for high temperature resistive heating. This process only further veried that temperature is dependent on the square of the voltage applied or directly proportional to the applied power. Temperatures near 350o C were achieved, but at the cost of more than 50W of power. With such a large power requirement, 32

the ability to store and then supply high peak power will be crucial for in-space rigidization via this method. However, the total electrical energy may oset high power requirements if the material can be cured in a short amount of time. The novel thermoset resin, UnyteSet 201, used in this study also helps to diminish the power needed for high temperature cure. This material cures at temperatures in the range of 150 200o C, whereas a resin such as PETU cures at more than 250o C. If the temperature-power relationship shown in Figure 2.15 is referenced, the power to reach the cure temperature can be reduced by more than 37% by using the UnyteSet resin ( 22W) instead of the PETU thermoset ( 35W).

2.4

Active Temperature Control of Resistive Heating

Open-loop resistive heating, which has been addressed up until now, applied a given voltage to a resistive material and the resulting temperature increase was measured. However, in order to achieve a desired temperature through voltage selection, a model for the system must be well-dened. This type of control, called open-loop control, applies a controller, independent of the output, to the plant. In eect, the appropriate control signal is rst computed based on a desired reference temperature. The control signal, which is then applied to the actual system, generates the desired temperature response. The possibility of open-loop control as an eective control strategy is investigated for this system.
Disturbance Input

Reference Input

Controller

Control Input

System (Controlled Object)

Controlled Output

Figure 2.16: Open-loop control scheme [9].

2.4.1

Model Formulation

For open-loop control to be eective, an accurate model of the controlled object must be known. In resistive heating, a model for a known length of the CFRP material can be developed as a function of applied voltage.

33

qc

, m, Cp i

qg
Ac z L -V

+V

Figure 2.17: Theoretical heating element used in developing a system model. The resistive element (Figure 2.17) of known dimensions (length, L, and crosssectional area, Ac ) and material parameters (resistivity, , and specic heat, Cp ) is subjected to an applied voltage, V which results in a current, i, owing through the sample. This model accounts for the heat generated, qg , by Joule-eect heating and the heat given o due to free-convection, qc . Further, it is assumed that heat conduction in the radial direction and thermal radiation from the sample are negligible. In doing so, we apply a lumped capacitance simplication, which assumes that the the temperature of the solid is spatially uniform at any instant during the transient process [33]. Simply, the lumped capacitance method negates conduction by assuming that temperature gradients within the solid are negligible. The Biot number, Bi , provides a measure of the temperature drop within a solid relative to the temperature dierence between the solid and its environment. It can be calculated to validate this assumption [33] Bi = hLc < 0.1, k (2.13)

where, h is the convection coecient between the solid and the surrounding air, Lc is the characteristic length (Lc is half of the radius for long cylinders [33]), and k is the thermal conductivity of the material. For carbon-ber, axial thermal conductivity can range anywhere from 5 200 W/mK [2]. However, the convection coecient, h, depends on temperature, uid properties (density, viscosity, thermal conductivity, and specic heat), surface geometry, and ow conditions and is more dicult to approximate. So, for now, let us assume that lumped capacitance can be applied to this system. The rst law of thermodynamics, which demands conservation of energy, is then applied to the resistive 34

element in E out = E. E (2.14)

For this process, the heat generated, qg , comes from the applied electrical signal, the energy given o is due to convection, qc , and the change in energy of the sample results in a temperature increase q g q c = mcp dT . dt (2.15)

Applying Joules law, and substituting in the expression for convection heat transfer, we obtain an expression for the heat balance of this system, dT V (t)2 hAs (T (t) T ) = mcp , R dt (2.16)

where V (t) is the voltage potential across the length of the sample, R is the eective resistance of the sample, As is the surface area, T is the ambient (or lm) temperature of the surrounding air, m is the mass of the sample, cp is the materials specic heat, and T is the temperature of the sample at any point along the length. Re-arranging this expression yields the following rst-order, linear, ordinary dierential equation: mcp V (t)2 dT + hAs T = + hAs T . dt R (2.17)

Written in standard form [36], Equation 2.17 is simplied to dT + pT (t) = f (t), dt where p= and f (t) = V (t)2 + hAs T R 1 mcp . (2.20) hAs mcp (2.19) (2.18)

The solution to the above dierential equation can be solved via many techniques, variation of parameters, undetermined coecients, etc. [36]. More importantly, it provides a relationship between the sample temperature, T , the material parameters, and the input voltage, V , T (t) = ept where again, p= hAs mcp (2.22) C C pt e + T p p , (2.21)

35

and C= V (t)2 + hAs T R 1 mcp . (2.23)

Comparing Equation 2.5 to Equations 2.21 and 2.22, and recalling Equation 2.8, the time constant for the temperature response, = mcp 1 = , p hAs (2.24)

is a function of the material properties of the sample (m and cp ), the sample dimensions (As ), and the convection coecient, h. Having a relationship between temperature and applied voltage allows for an openloop controller to calculate the exact voltage input based on a desired temperature. However, when an exact understanding of the system breaks down and accurate material parameters are not known, open-loop control suers. For the sake of seeing how closely the current model can control temperature, we can model the input power required to cause a desired temperature increase. It is important to note that the applied signal, V (t), is modeled as a unit step function multiplied by the magnitude of the voltage. The Laplace transform of this term can be written as L V21 V (t)2 V2 L [u(t)] = = R R R s (2.25)

where u(t) is a unit step input. For the entire solution, we start with the original dierential equation (Equation 2.17) and apply a Laplace Transformation (mcp ) [sT (s) T (0)] + (hAs )T (s) = V21 + (hAs )T . R s (2.26)

Grouping like terms and re-arranging yields an expression for the input power, V2 = s [mcp s + hAs ] T (s) mcp T (0) hAs T , R (2.27)

as a function of instantaneous temperature, T , initial temperature, T0 , and ambient temperature, T . This expression can then be written in block form and denes the open-loop controller applied to the system in Figure 2.16. Notice that both the ambient temperature, T , and initial temperature, T0 , are factors in the performance of this controller (Figure 2.18). If these values are the same and the temperature, T (s), is dened as the increase in temperature from its initial value, T (s) = (T (s) T (0)) , 36 (2.28)

CONTROLLER

T (s)

hAs mcps+hAs -

T (s)

mcps+hAs

V(s)2/R

T0(s)

mcp mcps+hAs

Figure 2.18: Block diagram representation of the open-loop temperature controller. then the transfer function, G(s), between temperature increase and applied power can be written G(s) = 1 . mcp s + hAs (2.29)

Equation 2.29 states that an increase in temperature for a step input voltage is rst order, represented by an exponential growth. The general shape of this rst-order model conforms to the measured exponential growth temperature responses (Figure 2.10). A change in ambient temperature, which can aect the system during resistive heating, can also be thought of as an external disturbance (Figure 2.19).
Ambient Temperature

Desired Temperature Initial Temperature

Controller

Input Power

System (Controlled Object)

Sample Temperature

Figure 2.19: Block diagram of temperature control, via an open-loop controller. If the ambient temperature is fed into the controller, as shown in the block diagram, then the controller can account for this instantaneous value when it computes the control signal. The controller, on the other hand, is completely independent from the actual sample temperature (controlled variable) and thus, in general, it cannot account for external disturbances to the system. An underlying concern with this technique, then, is whether 37

or not the designer has an accurate system model. In this case, our inability to accurately measure or calculate the convection coecient and various material properties (, m, cp , As , etc.) weakens the eectiveness of an open-loop controller. Nonetheless, predictive open-loop heating was attempted on CFRP samples.

2.4.2

Predictive Joule Heating Results

Using Equation 2.21, the temperature response was simulated for both the heating (V = 0) cooling regions (V = 0). Matlab was used to generate the predicted response for various voltage signals. Specically, the code allowed the user to specify the initial heating and nal cooling time periods as well as dene the input voltage value. The user also species a temperature window in which the sample was allowed to heat and cool for a selected number of cycles at the same voltage level. The code then computed when to turn on and o the voltage in order to mimic these temperature swings and used these times to generate the voltage signal pattern automatically. This signal was read directly into Simulink and an exact replica was sent out of dSpace into the material. In this manner, the predicted response to a unique voltage signal was directly compared with the measured response from the same input. By using a combination of estimated material parameters and the experimentally measured time constants, the terms mcp and hAs (shown in Figure 2.18 and Equation 2.27) were approximated for this system and implemented into the predictive response simulation.
65
8

60 55 50
6

Temperature - C

45 40 35 30 25 20 Heating (V is ON) Cooling (V is OFF) 0 100 200 300 400 Time - s 500 600 700

Voltage - V

0 During cycles: Voltage is OFF for: 4.9357s. Voltage is ON for: 98.5982s.

-2

-4 0 100 200 300 400 Time - s 500 600 700

Figure 2.20: Simulated temperature response for an input of 6V and a temperature window of 10o C (left). Corresponding voltage signal generated to cause the presribed heating (right). Attempts to model and simulate the temperature response of the system proved

38

olsample31.mat 60 55 50 45 40 35 30 25 20 Th1 Th4 Th6 Ambient Temp 65 60 55 50

Open Loop Sample 31 (3-24-05)

Temperature - C

Temperature - C

45 40 35 30 25 20

Measured Response

Heating (V is ON) Cooling (V is OFF)

100

200

300 Time - s

400

500

600

100

200

300 400 Time - s

500

600

700

Figure 2.21: Measured temperature response induced by the presribed voltage signal (left). Comparison of the simulated temperature and the average measured temperature (right). inaccurate. While the general shapes of the curves matched, the nominal temperature values diered signicantly (Figure 2.21). As a result, temperature control via an open-loop control does not provide accurate temperature response within the material. The inability to eectively model the material and correct for external disturbances further negates this approach.

2.4.3

Feedback Temperature Control

Feedback control, on the other hand, updates the control input based on discrepancies between a measured controlled variable and the desired reference input. This technique still requires a controller but does not require the system to be completely known. What feedback control does require is a way to compare a measured value (say, temperature) with a desired value throughout the controlled process. It is this step that typically introduces sensors into the picture. The experimental setup, described in Section 2.2, demonstrates
Disturbance Input

Reference Input

+ -

Error

Controller

Control Input

System (Controlled Object)

Controlled Output

Sensor

Figure 2.22: Typical feedback control structure [9].

39

how temperature was fed back into Simulink/dSpace such that the control algorithm could compare desired and measured temperatures. For the application of resistive heating temperature control, a PID controller was selected. The basis for this control strategy stems from the success of proportional-integralderivative (PID) controllers as being robust, yet simple. The performance of the controlled
Kp
+

Ki

1/s

+ +

Kd

Figure 2.23: Transfer function for a PID controller. system depends on the selection of three control gains: proportional gain (Kp ), integral gain (Ki ), and derivative gain (Kd ). The process of determining the best values for the respective gains has been addressed by many and described by VanDoren as often more of an art than a science [37]. Notably, Ziegler and Nichols developed techniques for tuning these control gains in order to match desired performance [38, 39]. An understanding of the three control gains and their functions determines how they may be adjusted to achieve the desired level of control. Proportional control, with gain Kp , outputs a control eort, u(t) = Kp e(t), (2.30)

directly proportional to the measured error, e(t), between the controlled variable and the set point [37, 40]. With greater error, the larger the control eort becomes to diminish this dierence in temperature. However, proportional controllers tend to settle on the wrong corrective error, leaving an oset between the process variable and the set point. Integral control, adjusted through the gain Ki , generates a control eort proportional to the the sum of all previous errors [37, 40]
tf

u(t) = Ki
t0

e(t)dt.

(2.31)

External disturbance cancellation is another benet of using integral control. The addition of the integral controller provides assurance against steady-state errors, but is not always 40

the end-all for corrective control. Many PI (proportional-integral) controllers respond quite slowly without the use derivative control, Kd . This type of control generates a control action proportional to the time derivative of the error signal, u(t) = Kd d (e(t)) , dt (2.32)

and is used to generate a large corrective eort immediately after a load change in order to eliminate the error as soon as possible [37]. A full PID controller, then, combines the three types of control and requires selection for all three gains. The corresponding control signal for this controller is then written as:
tf

u(t) = Kp e(t) + Ki
t0

e(t)dt + Kd

d (e(t)) . dt

(2.33)

If the ratio of control eort, u(t), to the measured error, e(t), is dened as the control sensitivity, G(t), then the Laplace Transformation of this equation can be taken to achieve the transfer function for the controller [41] G(s) = Kp + Ki + Kd s. s (2.34)

Shown in Figure 2.23, the block diagram form of this transfer function demonstrates that the summation of these control gains provides the overall control signal. Using PID, temperature control was applied to the resistive heating process. Specically, the three gains, Kp , Ki , and Kd , were varied and their eects on the system were experimentally observed.

2.4.4

Feedback Control Experimental Results

Using Simulink to construct and provide the control algorithm processing, and dSpace to record temperature measurements and output the control eort, feedback control was applied to the resistive heating process. The basic feedback structure shown in Figure 2.22 assumes that the dynamics of the plant include the material itself as well as the thermocouples and that the material temperature is consistent along the length of the sample. In eect, feedback sensor dynamics from the thermocouples were neglected and an average measured temperature was used as the controlled variable. The rst round of temperature control experiments involved applying a step input as the desired temperature signal. The ability for the measured temperature to then match 41

this set point was recorded. A typical response measured during these tests is shown and its dening characteristics are labeled.

36 34 32 30
Temperature - C

28 ts 26 24 22 20 18 0 50 Time - s 100 150 Avg. Final Temperature, Tf : 30.02 C Percent Overshoot, P.O.: 16.83% Settling Time, t s : 22.20 sec. Damping Ratio, zeta: 0.493 Proportional Gain, Kp: 0.5 Integral Gain, Ki: 0.1 Derivative Gain, Kd: 0

Figure 2.24: Representative temperature response taken during the implementation and tuning of a PID controller. Several aspects of this gure were used to evaluate the eects of the selected control gains, which are shown within the plotting area in Figure 2.24. First, the average nal temperature, or steady-state temperature, was recorded as a measure of accuracy. Also, the percent overshoot, a measure of how much the measured temperature (blue line) overshoots or goes beyond the desired temperature signal (red line), was recorded. The ability for the control system to reach a desired temperature without severely overshooting it is important when prescribing a desired temperature-versus-time curing prole. Lastly, the settling time, ts , of the measured response was calculated as the time that it took the measured (actual) temperature to reach and maintain a temperature within 5% (green boundary lines) of the desired temperature. In evaluating each of the controller gain settings, these pieces of information were reported and helped to establish the nal level of control. The results of this section are organized so as to reect the gain selection process that was performed. The eects of each gain are evaluated for step, ramp, and tracking desired temeperatures. The plots in Figure 2.25 demonstrate that the average temperature can be driven to match the desired temperature. However, temperature overshoot (43% and 15%, respec42

38 36 34 32

36 34 32 30

Temperature - C

30 28 26 24 22 20 18 0 50 Time - s 100 150

Temperature - C

28 ts 26 24 Avg. Final Temperature, Tf : 30.00 C 22 20 18 0 50 Time - s 100 150 Percent Overshoot, P.O.: 14.51% Settling Time, ts : 21.30 sec. Damping Ratio, zeta: 0.524 Proportional Gain, K p: 0.5 Integral Gain, K i: 0.1 Derivative Gain, K d: 0.1

t s K p: 0.2 Proportional Gain, Integral Gain, Ki: 0.1 Derivative Gain, K d: 0 Avg. Final Temperature, Tf : 29.97 C Percent Overshoot, P.O.: 43.16% Settling Time, ts : 81.60 sec. Damping Ratio, zeta: 0.258

Figure 2.25: The rst few attempts at selecting gains resulted in marginal control. tively) proves to be quite signicant. The rst plot shows that for a desired temperature of 30o C, the actual measured temperature reached 34o C at its maximum point. The measured temperature in the rst plot also experiences signicant oscillation and as a result takes almost 82 seconds to steady out.
45

34 32 30

40 t
s

Temperature - C

28 ts 26 24 22 20

Temperature - C

Proportional Gain, K p: 0.4 Integral Gain, K : 0.04 i Derivative Gain, K : 0 Avg. Final Temperature, T : 30.00 C
f

35

Proportional Gain, Kp: 0.4 Integral Gain, K : 0.04 i Derivative Gain, K : 0


d

30

Avg. Final Temperature, T : 40.08 C


f

Percent Overshoot, P.O.: 4.15% Settling Time, t s : 19.80 sec. Damping Ratio, zeta: 0.712

Percent Overshoot, P.O.: 5.91% Settling Time, t s : 31.50 sec. 25 Damping Ratio, zeta: 0.669

20 18 0 50 Time - s 100 150 0 50 Time - s 100 150

Figure 2.26: Through a trial-and-error process, control gains that decreased the overshoot (left) were chosen. Some temperature nonlinearities appear when the same gains are used for a higher desired temperature (right). The process of selecting the gains that drive the measured temperature to the desired temperature set point with minimal overshoot explored many gain values. Figure 2.26 demonstrates that with control gains of 0.4, 0.04, and 0 for Kp , Ki , and Kd respectively, causes the actual temperature to reach and maintain the desired temperature. Overshoot for this setting was diminished greatly and the average steady-state temperature accurately 43

matched a desired temperature of 30o C. These same control gains, when used to match a step input of 40o C, were not as eective. Temperature nonlinearities in the plant resulted in greater overshoot and longer settling time even though the nal temperature matched the desired temperature. In order to further verify this nonlinearity, the set temperature was increased even further and the temperature responses were measured for constant control gains. The controlled temperatures show higher amounts of percent overshoot of increased desired temperature values.
80

PO = 4.31% 70

60

PO = 3.56% Tset=35C Tset=45C Tset=55C Tset=65C Tset=75C

Temperature - C

PO = 2.91% 50

PO = 0.57% 40

30

PO = 0% * Kp = 0.40, Ki = 0.04, Kd = 0.04

20

20

40

60 Time - s

80

100

120

Figure 2.27: Temperature dependent nonlinearities within the plant result in dierent responses. A second study investigated the individual eects of each control gain on the resulting temperature response. For these tests, all but one variable were left constant. First, the proportional gain, Kp , was varied from 0.5 to 1.0 and Ki and Kd were held at 0 for a desired temperature of 30o C. Proportional gain, Kp , is a control term that scales the output control eort proportionally to the measured temperature error. As a result, this control gain sets the initial slope of the temperature response. For higher desired temperatures and thus larger initial errors, the proportional gain results in a steeper temperature increase (heating rate). Since Ki was held at zero throughout this study, each temperature response exhibited steady-state error and none matched the temperature set point of 30o C. With the proportional gain aecting the initial temperature rate increase, the integral gain was now varied from 0.025 to 0.100 while holding Kp , Kd , and the set temperature constant. As discussed before, integral gain is used to minimize the steady-state error of a controlled variable. Figure 2.29 demonstrates this concept for the temperature response controlled with a non-zero Ki value (red line). The observed eect that Ki had on the 44

29 28 27 26 25 24 23 22 21 Kp=0.5 Kp=0.6 Kp=0.7 Kp=0.8 Kp=1.0

Temperature - C

Set Temperature: 30 C Ki = Kd = 0

50 Time - s

100

150

Figure 2.28: Temperature response versus a varied proporional gain, Kp . temperature response was reected in terms of the controlled systems damping. This control gain, when relatively larger ( 0.1), caused the measured temperature to overshoot the desired temperature and then oscillate before steadying out. As the integral gain was reduced to smaller values, the controlled temperature experienced less overshoot, exhibiting more of a damped response. So, in order to minimize overshoot, which is a goal for matching the prescribed temperature response, the integral gain was selected to be relative small compared to the proportional gain (Kp = 0.5). Lastly, the derivative gain, Kd was varied for a constant temperature step input at xed values of proportional and integral gains. The derivative gain, according to literature [37], is designed to speed up the response (or reduce its settling time). However, the measured temperature response was not noticeably aected by the derivative gain at low levels. The selection of the Kp and Ki values from previous analysis have produced an accurate, highly damped temperature response. Adding derivative gain to the PI controller did not quicken the temperature response for the system. Even more, in attempts to reduce the power required to control the temperature, eliminating one control gain potentially lessens the magnitude of the control signal and thus reduces power applied to the material. For step response, it was discovered that a proportional-integral (PI) controller resulted in a temperature response that matched the desired temperature and minimized overshoot. However, this type of desired signal is not practical when it comes to prescribing a temperature versus time cure schedule [42]. In this case, it is more common to ramp (increase linearly) the temperature up to the curing temperature, hold it there, and then allow

45

2.5

32

30 2 28

Nearly identical slopes!

Initial Slope - C/s

1.5

Temperature - C

26

24 1 22

Set Temperature: 30 C Kd = 0, Kp = 0.5

Ki = 0 Ki = 0.1

0.5 0.4

0.5

0.6

0.7 0.8 0.9 PID Proportional Gain, Kp

1.1

20

50 Time - s

100

150

Figure 2.29: By increasing Ki to a value of 0.05, the steady-state error disappears, but the initial slope remains (left). Calculated initial slopes as a function of increasing proportional gain solidies its eect on the temperature response (right).
50

45

40

Temperature - C

35

Proportional Gain, Kp: 0.4 Derivative Gain, Kd: 0 Set Temperature: 40 C

Ki=0.040 Ki=0.035 Ki=0.030 Ki=0.025

30

25

Ki=0.100

20

10

20

30

40

50 Time - s

60

70

80

90

100

Figure 2.30: Integral gain eects on controlled temperature response. it to cool. With that said, the ability to track the desired temperature becomes important for evaluating the eectiveness of this controller. By applying a ramping desired temperature, the eects of Kp and Ki on the ability to track a changing reference temperature were examined. For these tests, Kd was held at 0 since it provided no additional control advantage during the step tests. The plots in Figure 2.32 provide several insights into the tracking ability of a PI controller. First, it is observed that the proportional gain had little eect on the controllers tracking ability. For each gain value, the controller lagged the desired temperature signal. Secondly, varying the integral gain resulted in a more noticeable eect. As the value of

46

45

40

Temperature - C

35

Set Temperature: 40 C Proportional Gain: 0.4 Integral Gain: 0.03

30

25

Kd=0 Kd=0.02 Kd=0.04

20

50 Time - s

100

150

Figure 2.31: The measured temperature response demonstrated a weaker dependence on the value of the derivative gain at low levels of Kd .
40 38 34 36 34 32 Ki = 0.03 Kd = 0 Tset Kp=0.4 Kp=0.5 Kp=0.6 Kp=0.8 Kp=1.0 36

Temperature - C

32 30 28 26 24 22 20

Temperature - C

30

Kp = 0.5 Kd = 0 Tset Ki=0.05 Ki=0.09 Ki=0.13 Ki=0.17 Ki=0.20

28

26

24

50

100

150 Time - s

200

250

300

22

50

100

150 Time - s

200

250

300

Figure 2.32: Ramping temperature responses measured for varying proportional gains (left) and integral gains (right). Ki increased, the controlled temperature overshot the desired signal at the end of the ramping section and exhibited more oscillation during the nal zero-order temperature hold. The values of Ki tested, which were larger than the integral gain (0.03) used in the rst plot, demonstrated that increased integral gain also decreased the lag error during the temperature ramp. Overall, selecting Ki for a tracking signal reveals the trade-o between rst-order accuracy for larger values of integral gain and zero-order accuracy for smaller values of Ki . This decision is simplied since some error can be tolerated during the heating (temperature ramping) phase of a cure schedule. More important, though, is minimizing the overshoot of the measured temperature and providing accurate steadystate temperature holding. By selecting an integral gain to be small, but non-zero, these 47

stipulations were met and the control signal, u(t) was kept to a minimum for given values of Kp and Kd . Using the selected values to mimic a full-cure schedule, the eectiveness of the controller is shown. The gain values used during this last demonstration were again varied to see the eect on a desired temperature prole that combines step, ramp, and hold patterns. The error between the desired temperature and the feedback temperature was also examined. Each combination provided a controlled temperature within a few degrees of the
50
1.5 1 Maximum: 1.025C Minimum: -0.671C Average: 0.030C

Error - C

45

0.5 0 -0.5 -1 50 100 150 200 250 300 350 400 450

40

Temperature - C

35
6

30

25

Tset Kp=0.5,Ki=0.1,Kd=0.04 Kp=0.5,Ki=0.08,Kd=0.04 Kp=0.5,Ki=0.1,Kd=0

Maximum: 4.102% Minimum: -2.331%

Average: 0.091%

Error - %

2 0 -2

20

50

100

150

200 250 Time - s

300

350

400

-4 50

100

150

200

250 Time - s

300

350

400

450

Figure 2.33: PI controller used to control temperature through a full, curingschedule-type prole (left). The associated error for the third combination of controller gains as a function of time (right). desired temperature throughout the test. Using the third combination, which employs only proportional and integral gains (PI), it was shown that the controller was accurate to 1o C (ignoring the initial temperature step). So, through an extensive experimental parametric study of the control gains, the nal Kp , Ki , and Kd values chosen are 0.4, 0.04, and 0, respectively. Without derivative control, but stemming from a PID control strategy basis, the resulting feedback controller is simply a proportional-integral (PI) controller. Results from the feedback controller experiments were expanded to higher temperatures by using the Xantrex XHR 300V-3.5A DC Power Supply/Amplier. Specically, it was desired to see if temperature dependent non-linearities in the system aected the PI controllers performance at more-realistic curing temperatures. Again, both step and tracking desired temperature signals were used as references. Looking at the temperature response for the desired temperature step input (Figure 2.34), it is evident that the overshoot increased slightly for higher set points. In addition to the average measured tem48

100 90 80 70

60 Run 1: Tset=45C Run 2: Tset=65C Run 3: Tset=85C

50

40

Energy values for each run: Run 1: E = 0.075452 W-hr Run 2: E = 0.14892 W-hr Run 3: E = 0.23058 W-hr

Temperature - C

60 50 40 30 20 10 0 0

Power - W
0 10 20 30 40 50 Time - s 60 70 80 90 100

30

20

10

10

20

30

40

50 Time - s

60

70

80

90

100

Figure 2.34: Temperature response measured for higher temperatures (left). Associated electrical power and energy for each test (right). peratures (black lines), the maximum and minimum temperatures are plotted. These values represent the variation in temperature measured by each thermocouple. Also, the electrical power and energy (the area under the power versus time curve) supplied to the samples were measured. For a step input of 85o C, the maximum power supplied was over 50W, though the total energy during the test was less than 0.24W-hr. This plot demonstrates that rigidization through resistive heating demands large amounts of power, but if cured quickly, uses only moderate amounts of energy. Again, the average measured temperature was used as the feedback control signal, and temperature dierence band between the maximum and minimum recorded temperatures is shown. While the average temperature was controlled to precisely match the desired

200

Temperature - C

150 100 50 0 0 50 100 150 Meas. Temp. Band Desired Temp. Average Temp. 200 250 Time - s 300 350 400

30

Energy Consumed (area): 1.6301 W-hr

Power - W

20

10

50

100

150

200 250 Time - s

300

350

400

Figure 2.35: Tracking ability of the PI controller for a high temperature curingtype schedule. 49

temperature signal, the gap between maximum and minimum temperature widened as a function of temperature. Further, the electrical power data collected shows that power levels decreased by ramping the desired temperature. Intuitively, this makes sense. For a large step input, the controller must output a large corrective signal to make up for a great initial error measurement. In terms of the chosen PI controller gains, these values provided accurate control even at the much higher curing temperatures. Discussed later in Chapter 3, the controlled resistive heating of carbon-ber reinforced polymers is used to induce thermoset resin curing and matrix consolidation, which rigidizes the material. As an example of a heating schedule used to rigidize the samples, Figures 2.36 and 2.37 look at the temperature prole, the electrical energy required, and the error in the controlled temperature. This test forecasts the ability of temperature-controlled resistive heating to match a desired temperature schedule. This sample was heated up 160o C
180 160 140 120 100 80 60 5 40 20 0 Desired Actual 20 Energy Consumed: 1.27 W-hr

Temperature - C

15

Power - W
0 50 100 150 200 250 Time - s 300 350 400 450

10

50

100

150

200 250 Time - s

300

350

400

Figure 2.36: Controlled temperature (left) and resulting electrical power (right) measured for the cure of CFRP sample. at a constant heating rate of 60o C/min and held at the curing temperature for 2 minutes. The heating process required 1.27W-hr of electrical energy over a total time of 6 minutes and 40 seconds, with a peak power of roughly 20W. During this controlled heating process, the materials temperature was kept within 8o C of the desired temperature at all times. Even more, during the dwell time at the curing temperature, the controlled temperature was within 2o C of the desired 160o C.

50

180 160 140 Desired Actual

8 6 4 2

Temperature - C

120 100 80 60
Error - C

0 -2 -4 -6 -8

40 20

-10 -12

50

100

150

200 250 Time - s

300

350

400

450

50

100

150

200 250 Time - s

300

350

400

450

Figure 2.37: Temperature prole (left) and dynamic error (right) measured during an actual curing schedule.

2.5

Infrared Thermography: A Visual Approach

In establishing the temperature measurement technique used to monitor sample temperature during the resistive heating process, it was noticed that each of the three thermocouples measured dierent values (Figures 2.9 and Figures 2.34 and 2.35). Were the thermocouples poorly placed (Figure 2.3), or does a real temperature gradient exist along the length of the material? Up until now, the three thermocouple temperatures were averaged at every point in time during the tests. This technique was chosen merely to simplify the data through the assumption that the temperature along the length of the carbon-ber samples was constant. After all, current owing through a uniformly resistive element should produce the same temperature at each point. In eorts to validate the evenness of the sample temperature, infrared (IR) imaging was performed. By being able to visualize how the material heats, it was hoped to better understand if and how temperature gradients are introduced into the samples.

2.5.1

Introduction to Infrared Imaging

Infrared waves, produced as thermal radiation from heated bodies, are electromagnetic waves whose wavelengths are beyond the visible wavelength spectrum [3]. The emission of thermal radiation occurs when the oscillations or transitions of the many electrons that constitute matter release energy. The temperature of the matter, a form of internal energy, sustains the electron oscillations and is thus related to the emission of thermal radiation

51

[33]. Infrared (IR) cameras detect thermal radiation by columnating the thermal radiation and focusing it onto a detector of known material.

Figure 2.38: Infrared detection scheme as found in an Inframetrics 760 IR Camera [10]. Infrared imaging, or thermography, is not new to the study of carbon-ber and other composites. Much work as been performed on using IR imaging to detect inhomogeneities that aect the performance of a composite. Specically, factors such as constituent concentrations (ber-resin ratio), orientation and distribution of reinforcement, voids, and matrixreinforcement bonding can be identied through IR. Further, this technique can also locate foreign material, ber breakage, and degradation [43]. A form of non-destructive evaluation or testing (NDE or NDT), thermography can also be used to quickly analyze composites without further damaging, or even contacting, the material. Jones and Berger performed thermographic inspection on glass-reinforce composite marine vessel hulls [44]. Favro, et al used high power photographic ash lamps to generate a heat pulse on the surface of berreinforced polymers and ceramics. Their work also included the study of crack propagation as identied through thermography [45]. Further, Sakagami and Ogura, of Osaka University, investigated the transient temperature distribution result from through-thickness and surface cracks in steel plates as well as delaminated CFRP composite samples [46]. In their study, they used Joule eect heating, through the application of an electric current, to induce thermal radiation emission from the samples. In the area of spacecraft, Tr` etout, et al researched the feasibility of applying thermographic NDT to the evaluation of satellite structures during assembly. The authors establish an experimental procedure for acquiring the thermal scans and present a method for processing the data. 52

In contrast to previous work [43, 44, 45, 46], the work performed in this document applies thermography in a dierent manner. Instead of using temperature distributions to detect inhomogeneities within the CFRP samples, the temperature distributions themselves will be used to quantiy the evenness of the heating process. Particularly, this study investigated the temperature distribution within the samples during resistive heating and included sample length, sample twist, and temperature as possible deterrents from uniform heating. Thermographic imaging was also used to conrm temperature values measured using thermocouples.

2.5.2

Thermographic Imaging Results and Discussion

An Inframetrics 760 Model IR Imaging Radiometer was used to capture still images of the samples temperature map during resistive heating. The application of power and measurement of temperature remain as they were used before, with the only additions to the setup being the IR camera and a television (for real-time viewing). It should be noted that the images obtained during this test were used qualitatively to investigate temperature distribution, and not quantitatively to look at nominal temperature values. Additional infrared thermography images of the resistive heating process were obtained using a Flir ThermaCAM EX320 infrared camera. This instrument provided increased accuracy in not only visualizing temperature gradients but also in measuring nominal temperature values. With this capability, it was desired to verify the temperatures measured by the thermocouples as well as detect if the thermocouples were aecting the temperature of the sample. The tests consisted of prescribing a temperature-time prole for the control system to match (using an average measured temperature as the feedback signal). During the heating process, the camera was focused on to the heated sample and still images were taken. The matrix of test conditions evaluated through IR imaging included sample length (and thus resistance), sample twist, and also temperature. If these factors contribute to non-uniform heating, it was hoped to notice this eect in the IR images. While this instrument can display infrared images in both black and white and color modes, the images will be presented in color. The black and white mode, while dicult to distinguish small temperature dierences in gray scale, was used to focus the camera on the object. Then, with the camera in color mode, the temperature gradient was monitored and recorded. Two 53

Hot
Th1 Th4
i

Th6

V+

VTelevision

IR Camera

Cold
Figure 2.39: Experimental setup for thermographic imaging of CFRP samples during resistive heating. drawbacks of this device were its rather poor resolution in color mode and its inability to record real-time videos. First, the eect that sample length had on the resistive heating process was examined. The measured temperatures were plotted versus the desired temperature and the electrical power and energy were also recorded. Increasing the length of the sample increases
Sample Length = 3" Sample Length = 4" Electric Power Consumed 6 Length=3" Length=4" Length=5" Length=6"

Temperature - C

Temperature - C

60 40 20 0

60

Power - W
0 100 150 Time - s Sample Length = 6" 50 200

40 20 0

4 2 0

100 150 Time - s Sample Length = 5"

50

200

20

40

60

100 120 140 Time - s Electrical Energy Consumed

80

160

180

200

Temperature - C

Temperature - C

Energy - W-hr

60 40 20 0

60 40 20 0

0.15 0.1 0.05 0 Linear Trendline: E = 0.020L + 0.019 W-hr

50

100 Time - s

150

200

50

100 Time - s

150

200

5 6 Length - in

Figure 2.40: Measured temperature responses (left) and associated electrical power supply (right). its resistance and mass. As a result, more power was required to achieve the same temperature and thus more energy was used during the heating process (Figure 2.40). However, infrared stills captured for these samples demonstrate even heating. As shown, the samples have a hot (dark red) core region running the length of the sample and the temperature

54

Sample Length: L = 3"

Sample Length: L = 4"

Sample Length: L = 5"

Sample Length: L = 6"

Figure 2.41: Still images captured for various lengths at a temperature of 60o C. cools radially outward. Secondly, the maximum temperature attained was thought to possibly induce and/or widen a temperature gradient. Four dierent maximum temperatures were used to obtain IR still images of the heating process. Again, the samples heated evenly along their lengths, with the only noticeably temperature gradient in the radial direction.
Maximum Temperature: Tmax = 60 C 150 Maximum Temperature: Tmax = 80 C 150 Maximum Temperature: Tmax = 60 C

Focusing in on a

Maximum Temperature: Tmax = 80 C

Temperature - C

100

Temperature - C
0 50

100

50

50

100 150 Time - s Maximum Temperature: Tmax = 100 C 150

0 50 100 150 Time - s Maximum Temperature: Tmax = 120 C 150

Maximum Temperature: Tmax = 100 C

Maximum Temperature: Tmax = 120 C

Temperature - C

100

Temperature - C
0 50 100 Time - s 150

100

50

50

50

100 Time - s

150

Figure 2.42: Controlled temperatures (left) and temperature gradients (right) via IR imaging of samples heated to dierent temperatures. segment of one of the samples allows for the radial temperature gradient to be seen more clearly. Heating of the thermocouples during this process was also noticed through the IR imaging. In addition, it was observed (Figure 2.43) during the cool down phase of the heating tests that the samples reach room temperature much sooner than the thermocouples. As a result, data collected during this phase was skewed by an articially high thermocouple 55

Figure 2.43: Up-close thermographic image of the CFRP material (horizontal) and thermocouple (vertical) during a controlled resistive heat. temperatures. This eect is not signicant since during the cooling phase, the material has already under-gone its curing process and is merely being brought back to a cooled state. The eect of sample twist on the temperature distribution was also studied. The initial resistance for the twisted samples was recorded prior to each heating test. Table 2.2: Sample resistance measured as a function of the number of axial twists.
Trial 1 2 3 4 # of Twists 0 5 10 20 Initial R 12.7 10.9 8.7 8.4 Image # 15 16 17 18 filename irtwist1 irtwist2 irtwist3 irtwist4

Increasing the number of twists did have a noticeable eect on the resistance of the sample (Table 2.2), which in turn reduced the amount of power required to heat the sample. The temperature along the length of the sample, though, was not aected by the twisting. This nd ensured that twisting can be used to help secure the thermocouples in place without introducing new thermal gradients. Further, as Figure 2.43 demonstrates, the hottest region of the sample was centered in the cross-section with a slight thermal gradient noticed in the radial direction. In general, thermographic imaging through the use of an IR camera has shown that the temperature is fairly constant along the length of the samples. The ability to measure the hottest core region of the sample at any point along its length determines how well the thermocouple measurements agree. Since poor thermocouple placement can 56

Electrical Power Consumed 8 0 Twists 5 Twists 10 Twists 20 Twists

Number of Twists: 0

Number of Twists: 5

Power - W

6 4 2 0 0 20 40 60 100 120 Time - s Sample Resistance 80 140 160

180 Number of Twists: 20

15

Resistance - Ohms

10

10 Twists - #

15

20

25

Figure 2.44: Electrical power and sample resistance (left) and thermographic imaging (right) of twisted samples. As the number of twists increased, the resistance and electrical power decreased, but temperature distribution remained even. result in false temperature readings being counted, an average measured value is not a good choice for the feedback control signal. When an averaged temperature is used, and one or more thermocouples is not accurately measuring the sample temperature, the average value is lowered. This results in the control signal increasing its output in order to correct for a low temperature and in turn, actually raising the true sample temperature above the desired value. Instead, using the maximum measured temperature (from any of the three thermocouples) more accurately represents the true core temperature and reduces the chance of incorporating false temperatures into the control algorithm. The net eect of using the maximum temperature as the feedback signal is to better represent the actual temperature within the sample, eliminate the risk of accidentally overshooting the setpoint, and minimize the control eort required to drive it. The actual temperature of a coated, carbon ber tow sample was validated during resistive heating using an Flir ThermaCAM EX320 IR camera. While this device technically measures heat ux, temperature values can be obtained for objects of known emissivity in a dened environment (i.e. ambient temperature). The emissivity of carbon ber, which was experimentally measured by Eto, et al [47], ranges between 0.90 and 1.00. The increased resolution of the images obtained with this IR camera provide clear evidence that the thermocouples are aecting the sample temperature. While the sample reaches an nearly constant temperature along its length between the two thermocouples, it

57

359.7C

SP02*: 171.5

SP01*: 357.5 SP03*: 165.2

20.6C

Figure 2.45: Temperature gradient obtained on a sample during resistive heating is drastically lower near the thermocouples. For the resistive heating schedule prescribed in Figure 2.45, the image was taken at a point when the thermocouples measured a maximum temperature of 160o C. The IR image veries that the temperatures measured through thermography are also near this value when in the vicinity of the thermocouples. However, away from the thermocouples the sample reaches a maximum temperature of roughly 360o C. The size of the thermocouples (20-gauge thermocouple wire is 0.81mm in diameter) in relation to the twisted tow size ( 1 2mm in diameter). The thermocouples, in eorts to place them securely and measure the maximum internal temperature of the tow, are lodged into the twisted bers. Because of their size, slight misplacements of the thermocouples result in drastically dierent temperatures through the thickness of the sample. Further, the large thermocouples disrupt the ber alignment and tow orientation in the near vicinity. This eect, which may cause changes in the ow of electrical current and/or introduce more surface area subject to convective cooling, results in decreased sample temperature at the location of the thermocouples. It is seen that for a temperature of 160o C measured by the thermocouples, IR thermography records a maximum temperature closer to 360o C. Selecting smaller, 36-gauge (0.13mm wire diameter) thermocouples provides a lessintrusive temperature measurement technique. Placing the thermocouples becomes easier and ber orientation in the tow is preserved. A resistive heating schedule that prescribes a maximum temperature of 160o C and 58

500 450 400 350 Desired 36 ga. J-type 20 ga. J-type

Temperature - C

300 250 200 150 100 50 0

50

100

150

200 250 Time - s

300

350

400

Figure 2.46: Temperatures measured by small and large thermocouples on the same sample diered drastically. feeds back the temperature measured by the original thermocouple accurately tracks the desired temperature. However, the smaller thermocouple measures a much higher sample temperature during this process. This smaller prole of this thermocouple is less intrusive on the sample and measures a maximum temperature of nearly 400o C. Comparing the two thermocouple readings and the IR image obtained for the same heating prole shows that the smaller thermocouple more accurately measures the maximum temperature of the sample. In both cases, the large thermocouple aected the sample temperature locally, signicantly underestimating the true temperature and causing the sample to overheat.

2.6

Conclusions

Through the use of feedback control, via an experimentally-tuned PI controller, temperature control of CFRP materials during resistive heating has been established. First, uncontrolled, open-loop heating tests were performed in order to observe the heating behavior of these materials as well as initiate a method for implementing resistive heating and temperature measurement. Temperature control, both in open-loop and closed-loop congurations were then addressed. Using a lumped-capacitance heating model of the composite sample, predictive Joule heating was performed with minimal success. In its place, a feedback control algorithm was eectively implemented; the controller stemming from PID control theory. In this structure, a feedback controller compares a measured temperature to a desired set point and adjusts its corrective control eort to minimize the dierence. Through many 59

experimental variations of PID control gains, proportional and integral gains of 0.4 and 0.04 were chosen, respectively. The tests revealed that the derivative control did not noticeably improve the controlled temperature response and was thus eliminated (Kd = 0). Further, the selected PI control was shown to accurately mimic a desired temperature prole for step, ramp, and arbitrary tracking patterns. Lastly, thermographic images from an infrared (IR) camera were used to visually observe the heating process and conrmed that the samples exhibit consistent temperatures along their length. The accuracy of measuring temperature with thermocouples was also observed through thermography, and it was determined that smaller thermocouples provided more accurate readings.

60

Vous aimerez peut-être aussi