Search This Blog

Wednesday 20 November 2019

Clean Jazz Guitar Amp Builder Notes — Part 2: Power Amplifier




This is part 2 of  a series about a complete prototype guitar amp.

Part 1 lies here
Part 2 lies here
Part 4 lies here


 Power Amp Stage Notes:


Section A   The Final Amp | 21 Watt power amplifier for clean jazz guitar.

Above — Schematic of the final power amplifier developed for this series. Much of the design keeps the VAS + VBE multiplier, and ultimately the final output pair from instant destruction when things go wrong. For example, a speaker disconnection, or accidental bench experiment electrical short.  Of course,  we likely can't provide fail proof protection to sustained adverse events.

This amp includes an abundance of medium power transistors: BD139 | BD140 , plus current limiting in the VAS + the final drivers.

Most resistors outside of the input differential pair circuitry are 1/2 watt rated.  Most caps are 50 VDC rated.  In the schematic above, I've got the 1000 µF filter capacitors going to signal ground for clarity. Actually, they return to their own special ground @ the power supply as discussed in Part 1.

Above — 12-bit FFT of the maximum clean signal output | 21 Watts. With a beefier power supply , ( such as +/- 30 to 40 VDC on the rails),  better heat sinks on the power follower, and slight adjustment of the 2 current source emitter resistors, this amp could work OK to <=50 Watts or so.

On the bench, I hooked it to my dummy load.  I applied a 1 KHz tone and drove it to the brink of distortion as described in here on my Dummy Load page. That's what you see in this FFT.




Above — PA board installed in the amp.


Above 2 photos — In chassis for testing.  I'm using some of the installed rear panel jacks as temporary paths for input into the PA. This allows PA testing and real-world tests of the preamp + tone shaping stages.  2 speaker output jacks wired in parallel provide output possibilities.

I'll go through some notes:

I build circuits from DC to ~ 1.5 GHz on FR4 board using Ugly Construction. Evidently, this frightens some of my readers.
 

Board Construction & Heat Sinking

I built the PA on 2-sided , 1 ounce copper clad board. Circuit paths, or islands were carved with a motorized tool while outdoors. I also isolated each  board mounting hole so the board does not get connected to chassis ground when bolted in. I seek star grounding for each circuit board.

The PA board gets lifted off the chassis by spacers that prevent shorting the power BJT mounting bolts to the chassis [ recall that they're connected to each collector terminal]. This also allows air flow under the PA board for cooling.


Above —Heat sinking the power followers. I bent ~ 20 gauge aluminum with a brake to fashion crude heat sinks. Further, under each heat sink lies a small piece of 24 gauge copper sheeting and under that lies more 24 gauge aluminum stock.

The copper sheeting + aluminum heat sink is secured to the main board with the main power transistor mounting bolt plus an additional 6-32 bolt below each transistor. The additional bolt presses the heat sink & sheets  firmly onto the main board  — and lowered BJT temperature 3-4 degrees C more than without the extra bolt during tests.

For my 21 W amp, this crude heat sinking worked OK for this, my first solid-state guitar amp.  In a test setup with a dummy load, I applied a 1 KHz signal to give 20 Watts output in the DSO & left it that way for 24 minutes ( a torture test ). The big NPN measured 47-49 degrees, while its PNP sister measured 54- 56 degrees C.  My heat sink resistors measured 82 degrees C . When playing guitar at a comfortably loud bedroom volume, the transistors run around 23-24 degrees C.

Transistors

For the small signal transistors, I call for the MPSA06 and MPSA56 BJTs, however, by this time, I'd run out of them and applied 2N4401/2N4403.  My experiments showed that the MPSA series transistors  are better matched from BJT to BJT and likely prove a better choice for the differential input pair at least.

While I chose the very husky MLJ21194 +  MLJ21193 as power followers, a 2SA1302  2SC3281 pair, or even the TIP 35C | 36C could work well.  All of my parts were brand name & purchased from reputable American dealers. I saw a couple of cheap bootleg transistors let their smoke out during my several PA builds under heavier current.

In previous versions of this amp,I employed T0-92 encased transistors for all but the finals and the BD139-140 BJTs in earlier versions of this amplifier. This proved a big mistake. With mishaps + during torture tests,  the VAS transistor quickly or progressively failed  and the VBE multiplier died instantly. This led to a current surge and caused the fuses on my AC transformer secondary side to blow.

While T0-92 transistors provided a lovely high Beta, they led to a lot of frustration. In audio amps, and in particular guitar amps, "overbuilding" seems important. Thus, I used a lot of medium power transistors — even for the current sources.

Current Sources

For my build, I ran out of BD140s, so I bolted two TIP42A's onto my circuit for the current sources transistors. I isolated their mounting points and also sought them for mechanical rigidity — a needed thing when you build using Ugly Construction. The current sources exhibited nearly 0 temperature drift once they ran for ~ 1 minute.

I closely measured them in a variety of situations and placed some of this data in Table 1 on the schematic.  The input pair get sourced with a lovely high impedance with a current that varies from 6.25 mA to 5.7 under test conditions. I sought around 10 mA for the VAS supply , and measured a range from 10.5 to 8.61 mA with the amp driven to a constant 3W, 5W , 10W and 20 W output. 

This shows what happens to my DC power supply under different amplifier power loads.  I placed in DC measures in Table 2 below:


Since I'm only going with  1 pair of finals, the VAS current needs to be high enough to fully drive the top half of the AC signal swing @ high output to a potentially low output impedance . For example, with two 8 Ω  speakers in parallel which doubles the current needed to drive a single 8 Ω speaker.  The emitter follower located before the VAS helps that cause immensely.

Ultimately, For lower current, distortion, heat sinking requirements and  DC ripple, I plan to stick to an 8 Ω load, however, i plan to experiment with different speakers and all of mine are currently  8 Ω speakers. If  I find 2 speakers I like better together than just a single 8 Ω unit, then I'll order these 2 speakers in 16 Ω for a final future amp.

Amplifier Topology

A basic 3 stage power amp that borrows heavily from the recommendations of Douglas Self. The classic differential transconductance pair serves as the input stage to subtract negative feedback from the input and drive an error voltage to the voltage-amplifier stage or VAS.

In professional power amps, input differential stage emitter + collector resistors get replaced with high Z current sources + mirrors.  Discrete transistor current mirrors improve the balance of the input pair across a wide frequency spectrum to cancel the 2nd harmonic. You'll see a variety of current mirrors used extensively within op-amps to boost linearity & reduce HF distortion.

Some designers employ boot-strapped resistors to supply the VAS current and this evidently can also work very well. I went with a current source.

The voltage-amplifier stage [ VAS ]  provides all the voltage gain in a solid state power amp. I applied Douglas Self's recommendation of driving the VAS with an emitter follower to reduce distortion and garner more current drive for the followers. 

The VAS circuit also contains 2 small COG or NP0 ceramic RF caps to stop HF oscillations. The current limiting transistor helps protects the VAS in event of a catastrophe.

When I first built the amp,  I did not connect the final power transistor collector nor base node wires. Every other part was wired in however. Thus I had a small functional amp to test and measure with a signal generator, DMM,  DSO or oscilloscope, plus a dummy load.


Above — Output of the amplifier with just the BD139-140 drivers in-situ with a +/-15 volt DC supply.  I first test any PA circuit with my +/- 15 VDC bench supply before going to higher DC voltages.

You won't get much power with the 100 Ω degenerative emitter resistors, but can fully test the amp without a lot of current flowing.  By tweaking the 1K pot on the VBE multiplier, I also set the bias across points a and b at 1.2 VDC with no input AC signal applied.  Thus the bias is at a good starting point for when you connect the power followers.

Finally, connect the power followers and adjust the bias while looking at the output in a scope or DSO plus FFT.  The quiescent current of the entire PA ran ~ 22.4 mA.  At full clean signal power it will consume around 714 mA if you connect an ammeter between the supply and 1 of the DC rails.

Of negative feedback, folklore and tribalism

This design applies mixed mode negative feedback: voltage and current negative feedback. Both provide series feedback to provide the same function --  to boost PA linearity.

Negative feedback also affects an amps gain, bandwidth and frequency response, plus its input and output impedance (Z). Negative current or voltage feedback affect linearity + gain similarly, but with respect to output impedance, current feedback increases output Z by the size of the feedback factor, while voltage feedback decreases output Z by the feedback factor. To calculate the feedback factor for each type, a different equation gets applied and I won't go into the math.

A classic HiFi solid state amp with just voltage feedback exhibits a very low output Z that is just fractions of an ohm.  Delivered power decreases as the speaker Z rises.

For decades, solid state guitar power amp designers have applied current plus voltage negative feedback. This leads some to suggest that the solid state amp is now behaving more like a transformer-coupled tube amp that offers a higher output Z  while exhibiting a lower dampening factor. 
Some engineers who design audio amplifiers hotly debate current versus voltage source drive. See this brief EDN article by Esa Meriläinen who argues for driving speakers with current rather than a voltage source. I've got papers by respected authors who state the opposite.  I'll stay out of this debate.

In my PA,  output current is sensed by measuring the voltage drop across a 0.22 Ω resistor in series with the output voice coil. This resistor senses a fraction of the output signal and feedbacks a small voltage to the input stage that is phase inverted and thus "negative feedback".  The voltage drop across the 0.22 Ω resistor goes up or down proportionally to the output current. This variable voltage is fed to the 510 Ω emitter resistor in the 2nd half of the input differential pair, As the fed back voltage increases, amplifier gain decreases. E.g. increased voice coil current increases the voltage drop across the current sense resistor and this bucks the amplifiers input signal.

The 10K resistor connecting the power follower collectors to 1/2 of the differential pair's base node provides the voltage feedback.

Output Impedance and Dampening Factor

Going back to your electronics training, for example -- studying linear power supplies.  In current feedback, the circuit attempts to keep the output current constant meaning that the output impedance has to be high.  For output voltage to be constant the output impedance has to be low. Regardless, whether constant voltage, or constant current, the power delivered is still a function of speaker impedance versus frequency.

Dampening factor is another term that may polarize some guitar amp lovers; especially by those who adore distortion. Dampening factor (aka dampening) = the ratio of loaded Z ( typically 16, 8, 4 Ω ) to the amplifier's output impedance. In guitar world, higher dampening is often considered a bad thing by the cognoscenti. A very low output Z solid state amp dampens the speaker more than a higher impedance transformer-coupled amplifier. This is true according to the arithmetic.

Raising the amplifiers output Z reduces dampening, however, the most significant factor affecting dampening is the series resistance of the speaker voice coil. I once significantly changed the dampening on a friend's solid state amplifier with no soldering — I put in a different speaker with a much lower voice coil DC resistance. He loved the result.

How I chose the 820 Ω feedback resistor.      Color you PA Tone

The other board bolted in the amp chassis thus far contains a split-supply voltage regulator for the op-amps used in the preamp and tone shaping circuits, plus this board also houses the first preamp & it's connected to a 1K gain pot.

I connected my guitar to the amp input and routed the output of the first preamp into the power amp input. It's interesting to listen to a guitar amp with no tone control circuit -- and just 1 volume control.

I connected an 8 Ω 12 inch Jensen speaker to the amp and played through it. Instead of the 820 ohm resistor, I place a 1K 2 Watt potentiometer in the feedback loop and tweaked the pot as I played.

I seemed to hear some changes in the bass & treble frequencies at different rotations of the pot. This aurally supports the claim that current feedback tends to boost lows and highs and thus provides a mid frequency scoop. Further, I seemed to prefer 1 particular setting on the pot.  I removed the pot and measured it with an ohm meter.

Continuing on, I tried 4 other 8 Ω speakers that I had on hand at the moment. I also tried different speakers in parallel.  In most cases, the setting on the pot that I seemed to prefer was different.  I also tried longer speaker wires  ( 16 foot, 8 foot versus 3 foot  ) -- this exhibited a small effect on 2 of the speakers but this in turn felt dependent on the pot setting.  Speaker cable resistance can potentially exhibit an effect on dampening in a very low output Z power amp.  Normally, I use a 3  foot ( 91 cm) speaker cable

The 1 pot setting that worked as a compromise setting was then determined: 811 Ω, I soldered in a 820 Ω 1 Watt resistor in for now. When I finally complete the amp and decide which speaker(s) to use, I'll repeat my test with a potentiometer.

The current feedback resistor allows for a potential change in the output impedance that can allow for some equalization of the speaker - power amp system in concert with the speakers frequency response + resonant frequency [ speaker resonance often lies somewhere between 75 - 120 Hz and varies from model to model ].  I wonder if adjusting the pot allows you to enhance the resonant frequency of the speaker?

This function paled in comparison to actually changing the speaker. Using speaker choice to dial in whatever desired guitar sound lurks in your head seems critical. Then you can try to tweak it with a current feedback resistor choice.




Section  B   Faux Amp


Above —  Schematic for a faux amplifier I built for plus and minus 15 volts all on 1 PC board to investigate PA stages, plus different preamplifer circuits.  Clean signal output = 9.5 Watts, as my bench split supply has a small transformer and normally gets used for developing small signal op-amp circuitry. This was my first split supply power amp and I learned a ton and gained some confidence.


Above — The finals were mounted directly on 2-side copper clad board. At under 10 W power, they did not get above 30C even during maximum power torture tests for 15-20 minute stints.


Above —  RCA inputs used for connecting the +/- 15 volt split DC supply to this board. This is OK for a low power experimenter board, or for op-amp circuitry, but not recommended for real power amplifiers.


 Above —  Entire board. In the forefront lies an RCA jack connected to the PA input.

Above —  FFT with maximum clean signal. Above this drive my power supply voltage sagged excessively and distortion increased exponentially.


Above — Time domain. Compared to above,  increasing the drive by 1 Vpp produces harsh clipping.



Above — I've connected a preamplifer module to the power amp. I'll cover several preamp schematics plus discussion in Part 3 of this series.


Above — Side view. You can see the power supply rails. Generally you want to keep them closer together, but this was my first go and gave me lots of building space.


Above — Output section. The emitters connect to their resistors through small holes. I wound the coil and held it together with tape. This OK for a temporary project. I connected the output jack to either my dummy load or speaker(s).

I enjoyed this little amp and it sounded great. Surprisingly, a 9.5W amp cranked up proves quite loud in an electronic lab.

Miscellany


Above — This power amp featured parallel power followers to reduce current-induced Beta drop when driving 4 Ω loads into 60 W. I learned from this board that the most important stage in your entire amp = the power supply.

To get desired a output power, your DC supply needs to provide the needed current without sagging too much under full load. Using transistors that are specified to have lower Beta drop, plus putting them in parallel does help the Beta drop issue. The BJTs = MLJ21194 & MLJ 21193. Heat sinking was provided by copper plating on the back side to the board.


Friday 1 November 2019

Signal Generator for Bat Receivers



Introduction


In summer months, around dusk, I venture into our garden to watch bats feast on flying insects. In my valley, bat research + conservation garners deserved attention.  Bats prove an important part of nocturnal insect control & also help cycle nutrients from wetlands into forests.
Bat detectors, or bat receivers input bat audio in the range of ~15 -120 kHz and output AF into head phones, and/or give a Fourier transform of the received waveform on a small screen, or, in some cases, also provide the bats vocalized frequency range.

Bats frequency modulate ultrasonic audio and rapidly sense the location of fixed and/or moving objects in 3-dimensional space as they fly.  I’ve watched bats sense and careen around my dipole antenna wires on many occasions. They amaze me.

Bat receivers seem to fall into 3 basic categories:

  • Direct conversion; also call heterodyne, or zero IF
  • Direct digital sampling
  • Down conversion using logic IC division to move the frequency into a spectrum humans can hear


I plan to design and build my own bat receiver and to be honest, this task seems fairly unremarkable for someone who on occasion, designs and builds radio equipment. What you might lack is test equipment.

In order to start, I sought, an ultrasonic signal generator to help me design front ends, test mixers, and so forth. I wanted a very temperature stable signal source that gave some outputs between 15 to 100 KHz.  I sought a reasonably low distortion sine wave signal into a well buffered output network with amplitude control.  I require a sine wave output to measure loss or gain in my filters, amplifiers and downconverters.

Main Circuit Schematics 
Above — Main oscillator with binary division + signal conditioning.

The 74HC4060 14-stage ripple counter seems delightful. It offers a number of flip-flop bit counters, plus an inverter so you can hang an oscillator off this IC.  For temperature stability and ease, I chose a crystal oscillator. I wanted an output of 200 KHz, so that gave many crystal choices — ultimately, I went with a 3.2 MHz crystal fashioned in the usual Pierce oscillator topology.  Selecting output Q3 with Pin 7 gave divide by 16 and a 200 KHz output.


Above — Inverter output of my 4060 clocked at 3.2 MHz.

The 74HC193 4-bit synchronous binary up or down counter takes the 200 KHz signal and further divides it according to the 4-bit WORD set by data input pins 9, 10, 1 and 15.  Pin 9 is soldered to ground and stays Low. Pins 10, 1 and 15 are preset High through the 22K resistors. Thus the function table shown in the schematic should technically have Low placed in front of the other bits. For example, divide by 7 = Low, High ,High, High.

Pins 10,1 and 15 are set low as required by turning on a MOSFET switch. The 2N7000 is a lovely, inexpensive way to ground lower current components. I measured the average ON resistance on my board at 3.2 Ω for my 2N7000 trio.

A front panel switch controls the data bits of the 74HC193. You could eliminate the front panel 7PST switch + FETS and just ply 3 SPST front panel switches, however, I prefer grounding data pins local to the counter IC  — and not having to remember switch combinations to set the bits.

Steering diodes keep the MOSFETs on or off as selected. Not all of them are technically needed, but they cost pennies and ensure stable performance.



Above — Output of divide by 2 from the 74HC193 synchronous 4-bit binary up/down counter.

Following the programmable counters, comes a 74HC74 D flip-flop. Nothing cleans up a digital pulse better than a flip-flop. As you switch through the 7 frequency channels, each output exhibits a 50% duty cycle and is well conditioned for low-pass filtration into a sine wave.

Onto the low-pass filters and buffer stage:


Above — My low pass filter banks + the AF gain control and output buffer.

All signal capacitors are polyester types such as  polyester film, or metallised film etc..
I employed the NE5532 op amp throughout.  The 1  µF input connects the signal to a single order low-pass pole set for around 100 KHz and then into a unity gain buffer.

All of the op-amps employ a gain of 1 which simplifies construction.  The signal then passes through a cascade of second order low-pass filters that employ a Q of just under 1 ; Sallen-Key topology; and a Chebyshev response with a maximum ripple of 1 dB.

All filters use the same capacitor values and you can see how changing the 2 resistor values affects the 3 dB response of each filter.  I chose 5 filter outputs. The 20 + 25 KHz filter and 14.3 + 16.6 KHz  get combined to save 1 op-amp + 4 capacitors + 5 resistors.  Since these frequencies are fairly close together, attenuation of the higher of the 2 frequencies is minimal. You could actually make the whole board with just 3 filters: 50, 20 and 14.3 KHz, however, attenuation of some frequencies could be a deal breaker dependent on your needs.

The worst filtration occurs at 100 KHz, where the 2nd harmonic is only 20 dB down. For me, that's OK since I don't plan to use this much and its still reasonable for measurement purposes.


Above — The 100 KHz signal in Time Domain.


Above — An FFT of the 20 KHz signal. The worst harmonic = 38 dB down from the fundamental tone.  In an FFT of the 14.3 output, the worst harmonic is 43 dB down. Good results from simple low Q,  low-pass filters with 10% parts.
 

Above — A Time Domain DSO capture of the 20 KHz signal.

The various filter outputs connect to a 5 pole switch ( I stock 7 pole switches so 2 lugs go unused).
I can select the appropriate low-pass filter and may even "double down" by using the next  lower frequency filter knowing that the signal will get attenuated.

A 1 K pot allows a range of outputs from down a maximum somewhere between 5 and 7 volts peak to peak down to  ~ 10 - 20 mV pk-pk.

The output passes through 1 more unity gain voltage follower then to an AC-coupled RCA jack.
I plan to drive circuits directly,  or connect a piezo-electric transducer, or a MEMS ultrasonic transducer (CMUT) as needed.

Photographs and Final Notes


Above — I designed this project on the bench and built it using Ugly Construction.  I added  a power on LED to the circuit.


Above — Photo of the 7 channel selector switch plus entire circuit board. I do not plan to put in a case. I'll use it like this over the next couple of years.


Above — Filter section knob, plus gain pot and output RCA jack. I'll trim off the unused copper board using aviation shears.


 Above — Unused filters designed for 16.6 and 25 KHz.