You are on page 1of 43

WHAT YOU ALWAYS WANTED TO KNOW ABOUT WAVE

SOLDERING BUT WERE AFRAID TO ASK

Author: Martin Tarr

Source:
http://www.ami.ac.uk/courses/topics/0225_wave/index.html
Wave soldering
The wave soldering process
Materials
Fluxing
Preheat
The solder wave
Aspects of practical machines
Maintaining performance
Some of this text is currently undergoing revision. Watch for updates!

Introduction
In this topic we aim to help you understand:
the principles of operation of a wave soldering machine
aspects of the materials used that affect the process
how machine parameters affect the final joint
how problems may be both solved and (preferably!) prevented.

This part is divided into three main sections: after a preliminary reminder about
the process sequence and the materials used, the second section describes the
three main stages of the wave solder process. The final section gives some
information about practical implementation, machine parameters and set-up,
but is deliberately less detailed than a process engineer would need – wave
soldering is a complex process, and there are many options and trade-offs.
Although we do not explicitly refer to other processes that use liquid solder,
such as lead tinning, wire stripping and Hot Air Solder Levelling for PCBs,
much of the information on materials, equipment and control contained here
will also be found relevant in those contexts.

The wave soldering process

Development of wave soldering

The advent of the printed wiring board made it much easier, quicker and
cheaper to assemble electronic equipment. The time saving benefit when
making multiple solder joints was found first with hand soldering. However,
bringing all the joints into a single plane, with the board as a barrier between
solder and components, also created a structure in which soldering could be
automated by solder dipping.

Exercise
Before reading further, think in some detail about your response to the question ‘What
are the potential problems in simply taking an assembly and dipping it in liquid solder?’.
In considering this, you should recollect any experience you may have had in hand
soldering (for example, reworking) or in tinning components by solder dipping. To jog
your memory:
Did you need to use flux to help create a joint?
What happened when the soldering iron came in contact with the flux?
What happened to the nice shiny surface of the solder on the bit after
just a few minutes exposure to the air?
Did you get a quickly wetted joint just by bringing iron and solder in
contact with the parts to be soldered, or did you need to move the
iron so as to scrub the joint gently with the solder and make sure all
the parts of the joint were exposed to liquid solder?
Did you experience any problems with spikes left on the joint because
of the way in which you moved the bit away whilst the solder was
still molten?
Our ideas on this can be found in the remainder of this section.

Probably you will have thought of most of the points below, which highlight
the challenges in developing any method of machine soldering:
Unless the surfaces are unusually clean, flux always has to be applied in
order to encourage wetting. When flux is heated, first its low-boiling
constituents evaporate, and then it starts to decompose, generating
smoke. It is easy to tin a component by dipping first in flux and then
in solder, because the vapour and fumes can escape easily. However,
just placing a flat fluxed board in contact with hot solder will trap
solvent vapour between the two surfaces, preventing even contact
between solder and joint, and resulting in solder skips
Within a short time of exposure to air, the surface of molten solder
grows a layer of oxide. Not only is this oxide unsolderable, inhibiting
wetting, but the layer impedes free flow of the solder
Unless the operation is very carefully carried out, it is difficult to avoid
leaving surplus solder or solder spikes, even when the joint is fully
molten when the source of solder is removed
Some degree of movement of solder relative to the surfaces to be
joined helps accelerate the wetting process, and is needed to make
sure that solder reaches areas of the joint that are difficult to access.
Solder spikes

Many of the first in-line machines used the ‘drag soldering’ principle, where the
board was moved across a static pot. The relative motion scrubbed the solder
across the board and allowed flux volatiles to escape, and solder peel-back from
the joint was enhanced by arranging for the board to exit smoothly at a slight
angle to the pot surface. Automatic machines fluxed the board before solder
immersion, and could incorporate pre-drying of the assembly to reduce the
quantity of flux volatiles. A cover layer of oil was generally used to reduce
oxidation, although this meant that cleaning was almost unavoidable.
Wave soldering, also known as ‘flow soldering’, was patented by Fry’s Metals in
1956 and by the mid 1960s had become commonly used as a way of enhancing
productivity and yield. Relative motion between board and solder is enhanced
by the movement of the solder wave, and the surface kept free of oxide by
drawing up fresh solder from underneath. By way of analogy, think how you
might be able to take (relatively) clean water from the body of a pond whose
surface is covered with algae!
Wave soldering is an in-line process, during which the underside of the board is
successively fluxed, preheated, immersed in liquid solder, and then cooled
(Figure 1).

Typical ‘straight-through’ style of wave soldering machine


Figure 1: The wave solder sequence

This process has to be carried out in a controlled and reproducible manner to


ensure a high yield of good quality joints at the lowest possible cost. As a result,
wave soldering machines have become increasingly sophisticated, in an attempt
to control the many variables.
The temperature experience of the wave soldered joint, the ‘profile’ (Figure 2),
typically shows a steady temperature rise to over 100°C, then a rapid rise to a
peak of 240–250°C at the time of solder immersion. On immersion, the area of
the board in contact with the wave rapidly attains thermal equilibrium with the
molten solder, so that all joints reach the same temperature. The fact that a
large quantity of liquid metal is present to transfer heat is a key difference
between wave soldering and reflow soldering, and one that explains the lack of
a stabilisation plateau region.

Figure 2: The wave solder profile


Measuring the temperature profile has been particularly important in wave
soldering to reduce the damage to surface mount devices: thermal damage is
less of a problem with through-hole components, because the board acts as a
heat shield.

Wave Rider’ used for measuring the temperature profile and


immersion time

As with reflow, there is a critical contact time for soldering to take place – that
is a minimum time a joint must be in contact with the solder to ensure a good
joint. This depends on the type of joint, the solder pot temperature, and the
board type – constructions differ in their thermal characteristics.
There is a corresponding optimum contact time for an assembly, which is just
long enough to ensure that all joints become fully wetted. This time will depend
on what joint types are in the assembly and must be as long as the longest
individual critical contact time. Generally contact times between 3–4s are
suitable for most applications, but 1–2s is used for boards with sensitive
components. Working backwards, contact time determines the required
conveyor speed and wave dimensions.

Where wave soldering is used

Wave soldering has numerous applications including component lead tinning,


component manufacture, hybrid circuit assembly, and continuous wire tinning,
but its main application is for soldering circuit board assemblies. The process
for a through-hole component starts with selecting the part, cropping and
forming it where necessary, inserting it into the board, and then applying
molten solder to form the bond between the circuit board and component
termination.
Through-hole components have to be held in position to prevent movement
during handling and soldering, and especially to prevent them being pushed out
of the hole during soldering. The upward force on the leads is a combination of
their buoyancy (leads are less dense than solder) and the pressure of the solder
wave. This ‘component lifting’ problem is most commonly seen with parts such
as connectors, with multiple terminations and often little interference between
the leads and the holes in the board. Mechanical retention may also be needed
where the leads are either to be left long or to be sheared very short before or
after soldering.

Lifted integrated circuit leads

The many ways of keeping components in place include:


'Clinching’ component leads after insertion, that is bending over the
end of the lead that projects from the board, so that the component
is pulled and held against the board
Preforming the leads, using the residual spring in the lead to ‘interfere’
with the hole
With dual-in-line packages, the leads on opposite sides are set at an
angle and have to be compressed to the vertical for insertion. Whilst
they interfere with the holes, and are thus loosely retained, the leads
on opposite corners are usually clinched for extra security
Fitting retention clips to the handling jigs
Applying temporary weights. These can be as simple as a bean bag, or
as sophisticated as a shaped and weighted magnetic cover that is
automatically loaded and removed
‘Shrink wrapping’ with a plastic film heat-shrunk over the entire
assembly. Problems arise when the components are not of a uniform
height: tall parts can become twisted, whilst nearby components are
left loose. Also the adhering plastic film may not allow flux fumes
and expanding air to escape, causing blow-holes and insufficient
solder rise
‘Spray webbing’: applying a top surface spray that glues the
components in place and comes off during cleaning. This can be
applied as a hot melt system or as an evaporating spray
Applying a molten low temperature material to the bottom of the board.
Various wax-like materials are used that become detached and float
to the top of the solder bath when passing through the wave.

The methods most frequently seen are the first four in the list above, but the
choice will depend very much on the design requirement and equipment
available. For example, although requirements for low joint profile are now
more normally met by SM solutions, the heat shrunk plastic film method is still
used for assemblies where the leads have to be cut short prior to soldering.
Whatever the method, careful attention must be paid to static protection for
sensitive assemblies. Also, the component leads must project below the board
sufficiently both to ensure contact with the solder and to create joints where
good wetting allows the underlying termination to be seen. This so-called ‘pin
witness’ forms part of the specification requirement for all through-hole joints:
if solder is just ‘plastered’ over the surface to cover the lead, as can happen if
the solder temperature is too low, there is no guarantee that a proper joint has
been formed underneath.

Bulbous solder fillets – no ‘pin witness’

Wave soldering surface mount components

SM components were originally conceived in the late 1960s for ‘hybrid


microcircuits’, using ceramic substrates. Parts were either hand-soldered or
reflowed using a hotplate. This second process presented no problem, as
circuits were usually of single-sided construction and the ceramic was stable,
with good heat conductivity.
However, soldering problems began when SM components started to be wave
soldered to polymer circuit boards. Whilst chip components presented no
problems, active component formats were not very ‘soldering-friendly’, small
outline integrated circuit packages (SOICs) and plastic leaded chip carriers
(PLCCs) being especially difficult to wave solder.
This is because the ends of the leads are too close to the relatively high body
mouldings. The solder wave finds it difficult to access these corners, because of
the high surface tension of the molten solder. Until wetting takes place, the
solder surface in contact with a component is like a balloon pressing against the
walls of a room – in a tight corner, at best it will only make contact at the
periphery.
This ‘angle of aspect’, formed between the upper edge of the component body
and the end of the solderable lead, is about 60° for SOICs and can reach 90°
for standard PLCCs.

Figure 3: The contours of SOs, PLCCs and QFPs

A similar sort of situation exists where SM parts are closely spaced, making it
difficult for the solder to access the joint. There is no single solution: as you will
be able to deduce from your later studies, this problem is both addressed during
board design and tackled during manufacture by using waves with high
turbulence and an appropriate angle of attack.
The trend towards thinner packages might be expected to make gull-wing
formats easier to solder, because the angle of aspect is reduced. However, the
reverse is the case: usually the lead pitch also becomes finer, and the reduced
separation increases the incidence of bridging. Table 1 indicates the range of
SM components for which wave soldering is suitable.

Lead pitch Applicability of wave soldering

1.27 mm (0.05 in.) straightforward


0.75 mm (0.03 in.) more difficult, requiring special layout provision

0.5 mm (0.02 in.) usually need to be soldered in an inert atmosphere

<0.5 mm (0.02 in.) wave soldering not recommended

Table 1: Range of application of wave soldering

‘View and touch-up’

The solder joint faults introduced by the wave soldering process are normally
‘major’ defects, that is they require rectification. Typical of these are bridges,
spikes and solder skips, excess solder defects being the most common. When
problems happen, they tend to affect wide areas of the board, and most test
systems have difficulty in dealing with a number of short-circuits on a single
assembly. It is therefore common practice to introduce a rework stage
immediately after wave soldering. If a cleaning process has been specified, this
rework is usually carried out after cleaning, but may sometimes be done
beforehand, provided that the board is not excessively contaminated by flux
and that care is taken to avoid the flux residues hardening and becoming more
difficult to remove.

Overall inspection by a trained operator


is a crucial part of process control
In what is graphically referred to as ‘view and touch-up’, an operator both
inspects and reworks the board. This is usually done under low magnification
using a magnifier. A benefit is that there is direct feedback from inspector to
rework operator on the location and type of defect! However, there are dangers
in assigning dedicated operators to this task:
It builds into the company ethos the belief that wave soldering is
intrinsically a low yield process, and that some element of rework is
essential, whereas good designs and well-controlled processes will
give good results without adding cost
Operators may attempt to produce a higher cosmetic standard of joint
than is actually needed, in the process actually reducing the reliability
of the joint by reworking it. This becomes more likely when the
operator load is reduced by yield improvements.

Wave soldering mixed technology assemblies

Mixed technology (‘Type 3’) assemblies are often made by combining reflow
and wave soldering. A typical process is to apply solder paste, place surface
mount (SM) components on the top of the board, dry the paste, and then
reflow the paste to create the joint. Next leaded components are inserted, the
board is inverted, adhesive is applied, SM components are placed, and the
adhesive is cured. After inverting the board once more, wave soldering
completes the process.
Note that the SM parts are totally immersed in the solder wave. To avoid being
washed away, they must be firmly secured to the board by applying glue before
placement and curing the adhesive afterwards. Typical glues used are epoxies
that are heat cured, the process taking place in a belt oven similar to that used
for reflow.
Unfortunately, with wave soldering it is not easy to incorporate complex
components on the underside, which limits the freedom of the layout designer.
It also introduces a second type of soldering process, one that is
(comparatively) low yield and difficult to control. The alternative of using solder
paste and reflowing all the components is attractive, and cost reduction and
process simplification are major driving forces behind what is called ‘intrusive
reflow’ or (more descriptively) ‘pin-in-paste’. We will be exploring that theme in
Alternative techniques and off-board assembly.

Materials

Flux

Fluxes used for wave soldering (also referred to as ‘preparation fluids’) are
usually low-viscosity liquids, and consist of a flux base combined with:
activators
solvents (‘thinners’)
surfactants and foaming agents (where needed).

There are four main categories of flux used in wave soldering:


Conventional rosin-based fluxes, which use a base of isopropyl
alcohol (‘isopropanol’ or ‘propan-2-ol’, but usually referred to as
‘IPA’) and contain 5–20% of ‘solids’ (rosin + activators). These
fluxes will leave residues, which may or may not need to be removed,
depending on their nature and properties
Similar low-solids fluxes that have a solids content in the range 1–5%.
They may be based on rosin or synthetic colophony substitutes.
Low-solids fluxes are designed not to require post-solder cleaning
Water-soluble fluxes, which contain active organic acid components
in an alcohol base. The residue from these fluxes needs to be
removed from the circuit assembly with a water wash
Water-based fluxes in which active acidic components are blended
with water. Depending on the concentration of active components in
the mixture, these fluxes may be classified either as ‘no-clean’ or
‘clean with water’ types.

The trend to no-clean and water-soluble fluxes resulted from the CFC ban
following the Montreal Protocol agreement of 1987. There have since been
environmentally driven moves both to control the discharge of water used for
washing and to reduce the emission of Volatile Organic Compounds (VOCs).
The result has been a trend towards using water-based no-clean fluxes. Because
water has a higher boiling point and is relatively slow to evaporate, these fluxes
typically have a higher (10–20%) solids content, and generally require
modification to the preheat part of the process. This aspect is discussed at
greater length in Design for eXcellence.
Flux changes in composition when in contact with the atmosphere. These
changes take two forms:
Loss of solvents, which increases both the solids content of the flux
and its viscosity. These parameters are normally deduced from
measurement of the specific gravity (‘SG’) of the flux, and solvent
loss can be compensated for by adding replacement solvent
(‘thinners’). However, with rosin-based fluxes used in foam fluxers,
the changes are very rapid, so constant monitoring and
replenishment is required
Oxidation of the flux, which reduces its fluxing effectiveness.
Nothing can be done about this, other than replacing it by new flux.
A typical recommendation is that flux exposed to the atmosphere
should be replaced after 40–50 hours of use.

In order to get consistent fluxing results, either the condition of the flux has to
be maintained scrupulously (with foam, wave or brush methods) or virgin flux
has to be used once only (with spray fluxing). Low solids and no-clean fluxes
tend to be most difficult to maintain in consistent condition, which in turn
means that the industry is moving towards sealed spray fluxing systems.

Solder

Although 60:40 tin : lead used to be favoured, the material of choice for all
current wave-soldering applications is 63:37 tin : lead, that is, Pb37A (in the
ANSI/J-STD-006 standard) or Sn63 (in QQ-S-571E). [The moves to lead-free
materials as a result of environmental pressures are considered in Design for
eXcellence]
As you will have learned, there are benefits of adding small amounts of silver to
tin-lead eutectic solder, and the Pb36A alloy with 2% silver is very commonly
used for solder paste. However, for wave soldering the benefits are normally
outweighed by the greatly increased cost.
Fortunately, using the weaker alloy is not a major problem: a wave soldered
joint tends to be stronger by design than a reflowed joint, being generally
applied to through-hole components and surface mount parts on coarse lead
pitch, with correspondingly large fillets. Restricting the use of Pb36A to solder
pastes, where the demands on the joint are more stringent, and the metal cost is
only a small percentage of the total, makes good economic and engineering
sense.
Note that there will always be a substantial differential between the quoted
trading prices of commodity metals and the cost of solders. Only a portion of
this represents profit, as there are real costs associated with the additional
refining and casting processes involved. Whilst specifications and
manufacturing processes vary, electronic solders are very much purer than their
counterparts in other industries, and have a lower oxide content.

Solder resist

Solder resist, or ‘solder mask’, is an organic coating applied selectively to the


surface of the board to restrict the contact area with the molten solder. As a
result only non-masked areas will be wetted. Solder resist is usually a permanent
coating and is left on the final product. It offers the benefits of increased
electrical insulation, fillet control, and greater simplification of circuit design.
Solder mask is essential for all surface mounting applications because of the
fine tracks and small pad areas.
The material selected for a resist depends on the final substrate (rigid or
flexible), the cost of the assembly, and the final product operating environment.
Solder resists are generally made of epoxies and polyurethanes, but other
materials such as acrylics and polyimides have been used.
It is extremely important that the chosen material forms a flat, smooth, hard
surface of consistent thickness onto the board, otherwise solder webbing will
occur. In addition, the materials must withstand soldering temperatures, as well
as exposure to flux and any cleaning solvents.

The main advantages of solder resists are that they:


reduce solder consumption
slow the rate of contamination of the solder bath
simplify printed wiring design by providing an insulating layer
can be used to control the shape of the solder fillet
permit closer spacing of traces covered by resist
provide a partial conformal coating to the end product
increase surface resistivity and reduce tracking in humid conditions.

This last item is particularly important from the reliability point of view – the
increase in surface insulation resistance (SIR) is around two orders of
magnitude.
The only downside for solder resist is that some resists may interact with solder
to produce unexpected effects: an example of this is solder balling which is
reported to take place during wave soldering under nitrogen.

Solder masks are mainly applied in one of three ways:


pattern printing (‘screened’)
wet photosensitive film (generally curtain-coated)
dry photosensitive film (hot laminated).
The latter two are necessary for fine-line applications because of their better
pattern definition, and liquid solder mask is the most commonly used for high-
density SM work, because of its thin coating and inherent consistency of
thickness. More details are given in Solder mask.

Self Assessment Questions


Make as complete a list as possible of the ways in which aspects of the materials used
might impact on the wave soldering process. Include in your list:
Both the plus points and any disadvantages
Some comments on how materials are selected
How these materials are used/applied.
Can you think of any aspects relating to the materials used that have not been considered
so far in this section? [If you have experience of wave soldering, you might think about
any materials-related defects you have seen]
show solution

Fluxing

The correct quantity of flux has to be applied evenly to the entire surface to be
soldered: unfluxed areas will not be soldered well, and over-thick deposits will
lead to voids and solder balling. It is also desirable to flux the inside of any
through holes, to aid wetting and solder pull-through. The in-line application of
flux can be achieved by a number of different methods, of which the main
current techniques of foam and spray fluxing are described below.

Foam fluxers

Until the mid-1990s, the foam fluxer was the type most commonly fitted in
automatic soldering lines. With this method (Figure 4) a flux ‘foam’ is generated
by passing low pressure air (<1bar) through aerator tubes immersed in a tank of
liquid flux. Despite their common name, most of these ‘stones’ are in fact made
of polypropylene! [Having said this, ceramic tubes are reported to be more
reliable and easier to keep clean] The tubes, often fitted in pairs, are designed to
break up the stream into tiny bubbles, and are covered by an open chimney that
channels the foam upwards. The assembled board travels across the crest of
this wave of foam, and a thin coating of flux is left on the board as the bubbles
burst.

Figure 4: Schematic of foam fluxer


The flux used needs to be specifically formulated, with additives to aid the
creation of a stable foam with small bubbles. Typically, the flux is of low
viscosity, achieved by having a relatively low solids content. The thinner may be
based on solvent or water, the latter becoming increasingly favoured because of
environmental issues associated with VOCs.
The normal foam fluxer without any special support is often labelled a free foam
head, and can reach a height of around 15 mm. If greater heights are needed, for
example, where long through-hole leads are used, a brush support can be
introduced on both sides of the foam chimney and the total depth can be
extended to above 25 mm.
The maximum head of foam for a given type of flux depends on the nozzle
type, the air pressure and the amount of flux above the aerator. The latter
affects the stability of each flux bubble and therefore the foam head support
and height. Figure 5 shows different types of nozzle used in a foam fluxing unit,
and illustrates both how contact width can be increased for a given flux type
and how the height of the foam head increases as the flux level is raised (3 > 2
> 1).
When setting up a fluxer, the flux level typically starts 10–15 mm above the
aerator, and is increased gradually (up to about 35–40 mm) until no further
increase in foam head height is observed. At the same time the bubbles become
smaller, because the foam head is better supported.

Figure 5: Different types of foam fluxer head


The amount of flux deposited by a foam fluxer cannot be controlled directly by
the operating parameters of the fluxer itself, but depends primarily on the speed
of the board set by the conveyer and the viscosity of the flux. Note that the
latter parameter depends on the solvent content of the flux and foam fluxers
need constant addition of thinner to replace that which evaporates.

Spray fluxers

Since the introduction of SM technology, spray fluxing has gained in popularity


because the fine droplets can be propelled into the narrow gaps between
closely-packed SM components, whereas foam bubbles may burst before they
reach the joints at the base of the component body.
Most spray systems use a pressurised canister of flux feeding a spray head at
which the required amount of material is atomised by air pressure or ultrasonic
energy. A major benefit of a sealed system is that the flux is applied as received,
so that there is no chance of contamination, and complicated flux controls are
unnecessary. There are other advantages that outweigh the increased cost of the
equipment compared to foam fluxers, and have made sealed spray systems
popular, even as a retrofit option:
the quantity of flux deposited per unit area of board can be adjusted
and is relatively well controlled
the area sprayed can be confined to the board, reducing flux usage and
avoiding flux build-up on any carriers
running costs are substantially reduced, because there is minimal
evaporation of flux thinner
by applying a thinner layer, better wetting is achieved and there is less
flux residue
long component wires (up to 18 mm) and high component density
present no problem
the technique works with a wide variety of fluxes.

Many of the spray fluxers in the market place have easily replaceable, air-
powered nozzles, often fitted in pairs. Heads are fixed to a moving bar that
provides a reciprocating action similar to that used in paint sprayers, in order to
give complete even coverage of the board (Figure 6).

Figure 6: Reciprocating spray fluxer

Combined with a board sensor, the spray fluxer can be set to spray only the
required area. Depending on whether the head movement is pneumatic or
motorised, the limits of head travel may be set mechanically or electronically.
The latter option makes it possible to control all fluxing functions by computer.

The main problems with reciprocating air-powered nozzles are that:


the size of the flux droplets is large and the volume and spray velocity
are difficult to control
the fine-aperture nozzles tend to become clogged
the head movement is unreliable and needs maintenance.

As a result, there are a number of competing spray systems that vary in the way
that they address these problems and achieve wide, even coverage. Many of
these have an ultrasonic spray mechanism, in which the energy in high
frequency sound waves atomises the flux.
The amount of flux deposited varies directly with conveyor speed, but can also
be controlled by adjusting the fluxer parameters. How this is done depends on
the equipment being used. Generally, the more sophisticated the equipment, the
more flexibility there is over the amount of flux deposited.
Common to all spray fluxers is the need for an efficient extraction system,
which gathers any over-spray and the aerosol of small flux droplets that
unavoidably forms wherever flux is atomised. This airborne flux must not be
allowed to settle on the top surface of boards or on the soldering machine at
large.

Removing excess flux

Flux application is rarely totally even, and a fluxed board will usually drip
surplus flux onto the preheater, reducing its efficiency and creating a fire
hazard. There are two ways of removing excess flux, both of which are usually
integrated into the fluxing unit:
A fixed brush the width of the machine, similar to those used in foam
fluxers for foam support
An adjustable angled air knife.
On current machines, the air knife is the option normally fitted, as this avoids
physical contact with the board. A brush head is only suitable for assemblies
that are not sensitive to such contact. An additional benefit is that the air knife
will drive flux up into plated holes to enhance top side wetting.

Self Assessment Questions


What are the main types of fluxer unit? Think about a range of different board types and
list the tasks for which each is suitable. Are there any problems or limitations?
Justify your choice of the method you believe to be best for a high-density surface
mount board?
show solution

Preheat

Reasons for preheating

A freshly fluxed board cannot be wave soldered successfully unless its


underside has been heated to a temperature of about 100°C, the exact
temperature depending on the flux being used. The reasons for preheating a
board are:
to evaporate volatiles
to start flux activation
to reduce thermal shock on the assembly
to meet the ‘thermal demand’ of the assembly.

Removing the volatiles from the under surface of the board


Should excess flux solvent be left on a board through insufficient preheat
application, a vapour blanket can form between the board and the solder wave.
Needless to say this not only slows down the heat transfer between the molten
solder and the board, but can also cause the solder to spit. This is a major cause
of small globules adhering to the underside of a board.
Where the flux thinners are solvents, they are easy to evaporate. However,
water-based fluxes are more difficult to handle since they can retain enough
moisture to spatter during the soldering process.
Removing volatiles requires removing saturated air from below the board. With
an inclined conveyor, there is usually sufficient movement caused by natural
convection above the preheater. However, with horizontal or near horizontal
conveyors, additional artificial air movement may need to be generated.
Even after preheating, a certain amount of ‘sizzling and spattering’ can usually
be anticipated when the board enters the wave. This is directly related to the
amount of volatiles still left in the flux. However, over-drying is not
recommended, as this makes the flux on the surface relatively immobile,
interfering with solder wetting because it is not easily displaced by the wave.
Note that preheating alone is not enough to remove volatiles (such as moisture)
that may have been absorbed into the PCB structure. These need to be
removed by pre-baking 1.
1 On soldering, any volatiles in the printed wiring laminate will give off gaseous materials and cause blow-holes and entrapped gas pockets in the

solder joint. This gas evolution may also create sufficient force to rupture the plated via barrel or cause delamination. These volatiles can be
contaminants deposited during storage, handling, and assembly operations, trapped processing solutions, organic volatiles from the materials used in
board fabrication, or natural moisture. However, pre-baking is a costly and time-consuming operation, and one which may impair solderability.
Users should only need to bake a board before wave soldering either when packages have been left open for an extended period or when blow-holes
are discovered.

Flux activation

Some fluxes like rosin, depend on heat (70–80°C) to become active at all. The
level of activity of other types of flux also increases with increasing temperature.

Reduction of thermal shock

The thermal gradient between room ambient and soldering temperature is


enough to cause serious damage to many materials, especially non-metals, and
SM components should be specified to be compatible with wave soldering. This
can be done simply by immersing the parts directly in a solder bath. Suitable
tests are referenced in IPC standard 9501 PCB Assembly Process Simulation for
Evaluation of Electronic Components.
For boards, the main concern is the distortion that occurs, often referred to as
‘warpage’ or ‘warping’. This is mainly caused by the difference in thermal
expansion between the underside of the hot board, which is exposed to solder,
and the cooler top. The result is that the centre depresses and pushes itself
further into the wave, whilst the sides curl up and may not come into contact
with the molten solder. The warpage is made worse by the random location of
holes drilled in the board, internal copper layers, and by the uneven distribution
of component weight.
Preheating the assembly reduces both the thermal gradient between the top and
the bottom of the board, so reducing the potential for warping, and the thermal
shock to which any SM components on the underside are subjected.

To meet the ‘thermal demand’ of the assembly

When making joints with liquid solder, you need to have sufficient thermal
energy available to ensure that the interface remains liquid. If you dip a totally
cold surface into solder, you will first freeze a film of solidified solder that
masks the surface but is not in intimate contact with it. This has the result that
heat transfer is impaired, so that the joint area takes longer to reach a
temperature high enough for the solder to wet. In all soldering processes, the
parts to be assembled must be hot enough before solder is applied to avoid this
happening.
In wave soldering, the heat required to raise the surfaces to the wetting
temperature comes from both preheating and contact with the solder wave.
These work together to supply the heat necessary for the joining process. The
higher the preheat temperature, the less heat is required from the wave. This
can be translated into a shorter time in the solder wave, or higher production
speeds.
Without an efficient preheating stage, high conveyor speeds would not be
possible, nor (during the brief time available) could the molten solder be
persuaded to rise through the plated holes in a multilayer board to form a
meniscus on the top surface.

Preheater types

Infrared preheater unit


Most systems provide the majority of their heat from underneath using infrared
radiation. For the first drying stage of preheat, panels are common. These are
hot plates or rods emitting long/medium wave infrared. For the more intense
heating required later, quartz lamps emitting short wave infrared are frequently
preferred for functional and maintenance reasons:
They are easy to control and give high heat output
They have an extremely fast response time, so can be switched on only
when needed, reducing the running cost
Fast response plus programmability makes them useful for systems
that are frequently moved from one product to another – there are
some machines can be programmed to change parameters between
individual boards
It is possible to protect the heaters with a glass shield, simplifying
cleaning
Unprotected lamps run at high temperature, so are self-cleaning
The reflectors behind quartz lamps are easily replaced when dirty.

Convection panels are becoming increasingly used, primarily to provide the air
flow that is necessary to dry water-based fluxes.

Convection panel in preheater


When preheating is applied only from below, the rate of temperature rise of the
assembly can be insufficient for heavy or densely populated boards, because too
much energy is conducted away from the underside. For that reason, many
machines have additional pre-heaters mounted above the work, especially in the
section immediately before the wave.
The preheat requirements of products vary greatly, depending on the thermal
demand of the assembly and the drying properties of the flux specified. For this
reason, many wave soldering machines have modular preheat systems, which
can be reconfigured on site for a specific application.
Whatever the set-up of preheaters in a machine it is most important that the
whole board receives the same amount of thermal energy, because uneven
preheating is a dangerous source of soldering faults. However most modern
soldering lines give a warning if a heater in the preheating section should fail;
some even prevent further boards from entering the machine in case of failure.
Boards still in the machine must of course continue to travel forward, if they
are not to be cooked or get stuck over the solder wave.

Self Assessment Questions


Before reading further, think about what happens during the preheating stage, and make
a list of the features you would look for in selecting a preheater design for a specific
application.
show solution

The solder wave

Construction

The solder wave provides direct contact between the solder and the component
joints on the PCB. It can be divided into two distinct physical events:
Final heat transfer – raising the surfaces of board and leads to wetting
temperature. This is a function of:
solder bath temperature
wave contact length
conveyor speed
wave dynamics

The supply of molten solder – providing solder for wetting and filling the
gaps. This is a function of:
the solderability of both surfaces
design (pad to component ratio and fillet control)
wave dynamic

As indicated, both parts of the process depend on wave dynamics or, more
plainly, on the shape of the solder being pumped, its fluidity, flow rate and
turbulence. There are a large number of wave designs but most are variations
on the same technology (Figure 7).

Figure 7: Working principle of a wave-soldering machine

Solder waves are produced by forcing molten solder upwards, from an area
where there is no dross, through a vertical conduit that ends in what is
commonly called the wave nozzle. This nozzle will contain baffles to ensure
uniformity of flow to the top. Originally, the wave nozzle had the form of a
narrow slot, set at right-angles to the direction of board travel, with the
emerging solder forming a hump of molten metal and falling in a symmetrical
wave over both sides back into the main solder bath.
The symmetrical wave was soon replaced by the asymmetrical wave shown
above, which gives neater joints, reduces solder bridging and permits higher
soldering speeds. The term ‘Lambda wave’, though an Electrovert trademark,
is often applied to this waveform because of its similarity in shape to the Greek
capital letter Λ.

Single asymmetrical wave


At the front of the wave, the solder flows in the opposite direction to the
board, providing a scrubbing action, and assisting wetting. At the back of the
wave, the solder flows in the same direction as the board, so that it peels away
smoothly, and as much solder as possible drains back into the pot, reducing the
incidence of shorts and bridges.
The reason for this effective action lies in fluid dynamics. The object is to drain
the solder excess from the pin, leaving only what can be retained by the
meniscus around the board/pin interface. Pins should therefore be moved
vertically away from the solder, since any other angle will expose a larger
surface area of the pin, and a greater force would be required to separate pin
and solder. This is why Electrovert chose to make the flow of the wave equal to
the conveyor speed, as this gives a vertical downward vector to the separation,
and hence the best drainage.
Depending on the wave design, the presence of the board may modify the flow
characteristics. On some machines solder does not move over the back plate
until the board comes into contact with the front of the wave, pushing the
solder over the back-plate, to leave a clean soldering surface and complete the
G shape.
With most types of wave solder machine, an impeller pump, driven by a
variable speed motor, propels the solder downward into a pressure chamber,
from which it flows up through a vertical conduit towards the wave nozzle.
This arrangement keeps the movement of the solder towards the weirs at both
sides of the nozzle as free as possible from any turbulence. Before the
introduction of SMDs, this waveform, with the board skimming through the
top of the solder wave, was the best way of achieving a clean, solder-bridge free
assembly at conveyor speeds of over two metres per minute.
To minimise dross, solder waves are generally only pumped when needed.
Sensors determine the imminent presence of a board, and the pump is then
activated, but allowing enough time for the wave to stabilise before contact
between board and wave occurs.
The double wave

Figure 8: The two pumps in a double-wave chip nozzle

Sources: Electrovert (LH); Iemme (RH)

The concept of the double wave is shown in Figure 8:


The primary or ‘chip’ wave is a symmetrical wave with an intentionally
turbulent wave crest. The high kinetic energy at the point where the
solder meets the board ensures that the solder finds its way to every
joint on the board
A secondary wave allows the solder to drain away from the board
without leaving behind any bridges or unwanted accumulation of
solder. This is usually a standard asymmetrical wave.
Close-up of turbulent chip wave
Most double-wave machines have two solder pumps, one for each wave, but
both take their solder from the same reservoir.
The chip wave leaves the joints somewhat untidy, with some bridging.
However, most of these imperfections are tidied up in the second wave, with its
smooth exit conditions.
A dual wave system may be used for both leaded and SM components, but on
most machines the pumps can be operated independently, allowing the chip
wave to be turned off for applications that have only leaded components.

Other wave types

There is an inevitable gap between the two waves, and this has been criticised
for several reasons:
A dip in joint temperature occurs between the waves, which may
partially solidify the joint, making it more difficult for the secondary
wave to perform its function
The time in contact with molten solder is extended
Especially with low-solids fluxes, the chip wave may wash away so
many of the flux residues that the board is unprotected during its
second exposure to solder.
Double wave system: note the extended distance between
the two waves
This has led to the development of a wave form that combines the functions of
both chip and main waves in a single asymmetrical wave. This is done by
generating a zone of high multi-directional kinetic energy within the wave at the
point where the board enters, typically using an electromagnetic transducer to
vibrate a movable element within the wave. The position of the element defines
the vibration zone, and its amplitude of movement can be varied to optimise
the action and eliminate defects. The exit side of the wave follows the normal
asymmetrical pattern with a smooth, laminar overflow. The process is claimed
to be reliable, repeatable and efficient.
In some machines, the vibrating wave is preceded by a conventional high-
energy chip wave – belt and braces! In defence of the manufacturers, it must be
said that all these wave functions can be selected, so that the machine can be
configured very flexibly for a wide variety of applications.

Interaction of wave and assembled board

The interaction of a conventional laminar wave and the assembled board can be
divided into three zones, as shown in Figure 9:

Figure 9: Schematic of wave showing the three zones


Zone 1 – Point of entry

The point of entry is at the most dynamic part of the wave since the directions
of board travel and solder flow are in direct opposition to one another. At this
point the solder flows rapidly down the wave, while the board moves in the
opposite direction.
This differential motion creates a washing action that removes the flux from the
board, and will also flush away organic layers such as any surface
contamination.
Flux removal is total on the metallic pads where wetting occurs. However,
some of the more viscous fluxing materials may cling to the laminate between
conductors. The extent of flux retention depends on the physical layout of the
lands and components, but inevitably some flux is left on the bottom of the
board between lands.

Zone 2 – Heat transfer and solder rise

If only the bottom of the PCB had to be wet, the wave solder operation would
be complete shortly after the point of entry. However, component leads with a
substantial heat content also need to be soldered, and plated through holes
must be wetted and filled with solder.
This filling is primarily produced by wetting taking place between solder and
metallisation and the resulting surface tension forces: capillary action ensures
that solder will rise up the holes. There is some contribution from wave
pressure, although, as the board approaches the point of exit, the upward push
due to fluid dynamics decreases in importance.
In addition to the heat supplied by the wave, heat is absorbed by the assembly
during the preheat stage. This additional heat is critical in good fillet formation.
A board that is already warm can pass more rapidly through the hot wave.
Contact time is thus reduced, which decreases thermal damage. Note that the
greatest damage to components and board happens during exposure to the
wave. This is because the damage mechanisms are accelerated at higher
temperatures, whereas the assembly can tolerate the lower temperatures of
preheat for an extended period with few ill effects.
Because of the importance of heat transfer, the part of the wave between the
point of solder to metal contact and the point of wave exit is often referred to
as the heat transfer zone.

Zone 3 – Point of exit (peel back)

In order to achieve the best wave solder results, a high degree of uniformity in
fillet configuration has to be obtained. This ultimately eases inspection and
dramatically reduces unnecessary operator touch-up. To obtain such uniformity
the forces shaping the underside of the fillet at the wave exit must be
controlled. These forces fall into two categories:

Surface energies. These forces are predictable and can be subdivided into:
The interfacial energy between the solder and the base metal. This is
affected by solderability but is independent of the environment
The cohesive force of the liquid solder, which is greatly affected by the
second phase with which it is in contact (flux, or air).

Hydraulic forces. These are often random and depend on the following
factors:
The quality/purity of the solder being pumped, and hence its fluidity
Any oxide layers on the solder surface
Aspects of the wave design, such as turbulence and the direction and
rate of flow
The angle of the conveyor
The configuration of the exposed metal areas on the board
The distribution of the thermal load.

The best point of exit corresponds to where these hydraulic forces can be
neutralised. This is achieved by withdrawing the fillet from the wave at a static
location, which is found where the board travel speed and direction are similar
to that of the solder flow.

Self Assessment Questions


Describe the way in which a wave acts to create solder joints, in particular what happens
in each part of the wave.
Which waveform do you believe would be best for a low density mixed technology
(Type II) assembly? Justify your choice!
show solution

Aspects of practical machines

Board handling

There are two methods of holding PCB assemblies in the conveyor:


Fixtures (or ‘pallets’, or ‘templates’, or ‘carriers’), which may be
adjustable or made specifically for a particular board profile (Figure
10)
Finger conveyors, which hold rectangular boards by the edges and are
adjustable only in width (Figure 11).
Figure 10: Soldering fixture

Source: Blundell (RH)

Figure 11: Finger conveyor


Sources: Seho (inset)

Pallets

On larger machines, pallets usually ride on chains, pulled along by their own
prongs: on smaller machines, a synthetic rubber belt provides friction drive to
the pallet. Some advantages are that pallets:
will hold odd and irregular profiles
allow different board sizes to be intermixed
are suitable for multiple boards
provide a robust jig for the hand insertion of components
can be used for retention fixturing
can be used for selective soldering
allow the board to be angled to optimise solder drainage
provide simple and rugged transport.

Pallets are, however, not very cost effective for medium to large volumes,
because a high labour element is required to load and unload the pallets. This is
especially true if finger conveyors are fitted, since the loading and unloading
operations must be carried out carefully to avoid damage to the fingers. Pallets
also interfere with the final cleaning process and must be unloaded before
degreasing or washing. However, the pallets themselves need to be cleaned
occasionally in order to avoid contaminating the fluxer. Also, when pallets are
recycled whilst still hot, they tend to depress the foam in the fluxer, resulting in
skips or uneven coverage.
Fixtures and pallets can be made in many different materials, but the service
conditions are harsh! Pallet material must:
be non-wettable with solder
be resistant to flux
be resistant to cleaning fluids (preferably non-absorbent)
not warp in heat
not take heat away from the board
be easy to fabricate (usually by NC milling).

Anodised aluminium, Teflon-coated steel, and titanium are all used, the last
being preferred because it has no surface coating and is less subject to damage.
However, although easier to clean, metal jigs lose out to a range of non-metallic
materials similar to FR-4 laminate, except for light items such as clips. Fixtures
made from composites are easier to handle in production because of their
lighter weight and lower thermal conductivity, and have less effect on the
process because they absorb and retain less heat.
Durostone® is one such composite. Made of epoxy reinforced with 65–80% of
glass fibre, it has a similar thermal coefficient of expansion to the board it
supports. It is robust and easy to machine, but will wear in contact with metallic
surfaces. For this reason pallets are often fitted with metal glides, usually made
of titanium.
Depending on the production volume, pallets may be customised for the circuit
(in which case they can incorporate anti-flood protection, solder masks for
selective soldering and clamps for component retention) or general purpose
styles that can be adjusted to fit differing sizes of board. Typically, the clamps
on these flexible types will integrate solder dams to prevent solder washing
round the side of the board.

Belt conveyor with universal board carrier

Note that pallets have to be cleaned regularly, because they always pick up flux,
which decomposes as it passes through the machine. This even happens when
spray fluxers are adjusted correctly, because inevitably there is some over-spray.
Not only does this make the jigs difficult to handle, but it will impact on both
the yield and the cleanliness of the final product. No-clean fluxes produce less
of a problem but, after many passes, almost any material will build up a film and
then solder sticks to the film, rather than the metal.
Finger conveyors

Finger conveyors are permanently attached to the machine. The fingers, which
come in contact with the solder, are always made from titanium, which is
untinnable with any type of flux. Most of the fingers fitted have a vee-groove,
as this restrains the board vertically, and prevents it being pushed up whilst in
contact with the wave. However, unless very carefully maintained and adjusted,
fingers with this design are prone to dropping boards. It is common, therefore,
to replace a proportion of the fingers (say 1 in 3) by fingers with an L-shape, to
provide a more secure platform for the board. This design of finger is also
easier to use in conjunction with pallets, where the upward solder pressure can
be overcome by the combined weight of assembly and pallet.

Finger conveyors have the following advantages:


low labour involvement
automatic pick up and delivery
no heat mass
no volume limitations.

Finger conveyors are very cost effective from small to large volumes and fit
more easily into in-line conveyerised manufacturing systems. However, they can
only handle boards with parallel edges and cannot deal with mixed width
batches. They are less suitable than carriers for use with boards that are less
than fully self-supporting – flexible boards may be dropped, or dip into the
wave crest, leading to flooding.
For problem-free operation, the fingers on the conveyor must be properly
adjusted and kept free of tacky residues. Cleaning the fingers helps hold the
board correctly and stops contamination of the flux station. Unlike pallet
cleaning, an automated mechanism for doing this can be fitted to the machine.
In adjusting the conveyor, attention must be paid to:
thermal expansion relief
board height control
accommodation of width tolerances
turbulence created in the wave
finger damage and maintenance.

Conveyor system

Conveyor configuration

The conveyor system is a key element in the construction of a wave soldering


machine. On most British and American machines, boards travel on one single
continuous conveyor from one end of the machine to the other. The slope of
the conveyor is used to make a bridge between the low level of the hand
assembly operation and a higher level, from which the cooling section
(frequently separate) returns the product to standard machine conveyor height.

Single conveyor on an Electrovert ‘Electra’ wave soldering machine

By contrast, other European machines frequently operate with the sections at


different angles. One constraint that this design introduces is great difficulty in
incorporating a finger conveyor, and most such machines use ‘rubber bands’ at
the edge, driving pallets holding the product. However, there is an advantage
that the speeds at fluxing, preheat and over the wave can be set independently.
For example, with high-mass products, the pre-heat section can run at a slower
speed than is used for the soldering operation.

‘Humpback’ style of inert reflow oven

Conveyor vertical control

Because the depth of immersion of every part of the board surface in the crest
of the wave affects the final soldering result, geometrical precision is required
where the board travels over the crest, or crests of the wave. The position of
every board in the vertical axis must be defined, in reference to both its
longitudinal edges, to ±0.3 mm. Any sideways tilt of the board relative to the
wave crest must be held within these limits. It is advisable that they should not
be exceeded because some unsteadiness of the wave, and warping and bowing
of the board itself, must also be accommodated.

Conveyor speed

The speed of the conveyor is a critical parameter in the wave soldering process.
The main considerations are:
The heat received by the board is inversely proportional to the speed
at which it travels through the preheat stage at a given setting of the
heater panels
The maximum practicable soldering speed of a wave machine is
governed not only by the ability of the solder wave to get the
necessary amount of heat into the board within the time available,
but also by the complexity of its pattern and the density of its
population of components
Multilayer boards with a high heat capacity must travel more slowly
than simple single-layer boards
Boards with closely packed devices and fine-pitch multi-leaded devices
must travel over the wave more slowly to give the solder the chance
to flow into the narrow gaps between adjacent devices, and to drain
away from the fine pattern of leads.

Board support

Boards entering the wave must always be kept flat, otherwise some areas may
not be properly in contact with the solder. Also, long components, such as
connectors, at right angles to the conveyor, may be mounted flush with the
board at the ends, but with unacceptable clearance in the middle. This may
result in unsatisfactory joints, will add to stresses, and will ‘freeze’ the board in
its non-flat state. Even worse, the board leading edge may dip under the wave
front, allowing solder to come over the top of the assembly. Such ‘flooding’ is
very difficult to rework.

Solder flood on top of a board assembly


Boards may warp when heated. But, even without such warping occurring,
heavy unsupported boards may flex under their own weight, and thin boards
may be too flexible. Where some possibility of board sag is anticipated, and
finger conveyors are fitted to the machine to be used, there are three ways in
which this can be prevented:
by applying stiffeners across the board width – these would usually be
titanium clips mounted on the more vulnerable leading edge of the
board
by using pallets, preferably with clamps to hold the board flat
by running a central support cable along the wave solder machine, as
with reflow ovens.

Support cables are usually thin multi-stranded stainless steel wire, and move at
the same speed as the conveyor. They are adjustable across the width of the
conveyor, so that their position can be arranged to coincide with unused areas
of the board, such as the fret between circuits on a multi-circuit panel.

Self Assessment Questions


Summarise the benefits and drawbacks of the options for transporting an assembly
through a wave soldering machine. Which of these would you recommend for:
a standard 1.6 mm thick board?
a 0.7 mm thick board?
an assembly with a number of areas that need to be screened during
wave soldering?
show solution

Hot air knives

Although not a cure-all for badly designed assemblies, hot air knives are often
recommended for problem circuits. These provide directed streams of hot air
to separate bridges before they solidify, using the fact that solder bridges
between pads are less stable than those bridges that form the joints between
pads/holes and leads. Of course, this can only be done when the solder is still
liquid, so the gas flow has to be applied to the board as close as possible to the
point where the board exits the wave.
Mixed results have been reported with older designs of air knives, which were
fixed and operated across the whole width of the board: it was an art to get
them set up correctly, as a result of which they were often not used.
More recently, methods of selective de-bridging have been developed,
responding to user pressure for higher yields even with designs where lack of
space prevents ideal layouts. These use carefully controlled streams of warm air,
whose velocity and angle of attack is ‘fine tuned’ to avoid disturbing the
desirable solder bridges that form the solder joints. Just enough gas is applied to
disturb the capillary/cohesive forces that maintain the unwanted bridge,
without reducing the amount of solder available to make a good joint. The
excess solder is forced back towards the wave, and falls back into the pot by
gravity.

Schematic of selective debridging nozzle

A good impression of the way that the solder is blown away from bridges is
given in this Seho video clip of a hot air debridging tool in operation. The gas
flow is directed only to the parts of the board that have been identified as
potential sites for bridging, leaving the remaining areas untouched. As de Klein
and Schouten comment: “most solder bridging can be more or less predicted,
unless the bridges are caused by a lack of flux activity . . . (when) . . . solder
bridging can be found randomly across the board and there is no cure other
than improving the fluxing process.”

Inert atmospheres

Rationale

In a normal atmosphere, molten solder quickly acquires a tough surface film of


mixed tin and lead oxides. As soon as the solder is moved or disturbed, the
oxide skin breaks and mixes with the solder underneath. The resultant mix of
oxides and clean solder is called ‘dross’. Because the process of wave soldering
involves moving solder around and letting it fall back into a bath of molten
solder from a height, the formation of dross is unavoidable unless measures are
taken to protect the surface from oxygen in the atmosphere.

Implementation

One way of protecting the surface is to remove oxygen from the surrounding
atmosphere, and since 1985 much work has been done on finding practical
ways of providing a nitrogen environment for wave solder machines.

View through tunnel lid of dual solder wave operating in nitrogen

The advantages of soldering in this inert atmosphere are:


a substantial reduction in dross
decrease in solder consumption
reduced machine maintenance
more consistent process yield.

In addition inert soldering gives improved solderability by improving wettability


even with the best of modern fluxes and can minimise or even eliminate the
need for post-solder cleaning.

Users report that:


Reduction in solder consumption has resulted in projected annual
savings nearly equal to the cost of retrofitting the wave solder
machine with a nitrogen-atmosphere system
By eliminating dross, the maintenance required to clean solder pot and
nozzle has been reduced from once every 40 operating hours to less
than once every 1400 hours
Solder wetting and through-hole wicking have been substantially
improved, with the virtual elimination of skips and bridging.

Although nitrogen is not a truly inert gas, it remains by far the most popular
option because of its ready availability and low price compared to other inert
gases. The best performance comes from machines that are inerted throughout
their length and have entrance and exit air-locks (Figure 12). Such machines can
easily provide an environment containing <50 ppm of oxygen.

General view of inerted wave soldering machine

Figure 12: Wave-soldering under nitrogen – a ‘hump-back’ configuration

However, machines with this special construction are relatively expensive both
to purchase and run. For economic reasons, probably a majority of users chose
one of the solutions in which nitrogen is provided only at the wave surface. Not
only is such a ‘nitrogen wave’ generally available as an retrofit add-on to older
machines, but it gives most of the benefits of the fully inerted machine:
The rate of oxidation increases with temperature, and the preheat areas
are relatively cool, so that the majority of the oxide is formed in the
vicinity of the wave
Substantial improvements in the rate of dross formation can be
achieved with quite modest levels of nitrogen injection, and most of
the problem can be eliminated by reducing the oxygen content only
to <1,000 ppm.

Designs vary considerably in the ways in which they both inject nitrogen and
define the inerted volume. Typically they use nitrogen sprays or diffuse nitrogen
through porous stones, and take advantage of the fact that the board creates a
seal over the wave, retaining the inert atmosphere.

Self Assessment Question


From the perspective of someone who had designed a board that gave problems during
wave soldering, what kind of improvements might you expect if you were told that your
new assembler’s wave-soldering machine was inerted and had a selective debridging
capability?
show solution

Maintaining performance
If you have seen a number of wave soldering machines in different companies,
you will almost certainly have come across some machines that are in less than
pristine condition! Partly this reflects the nature of the process, and the
difficulty of removing dross and dealing with flux maintenance in older
machines. However, with newer equipment, your observation may relate more
to the low level of expectation of operators and management, used to machines
carrying out less exacting work. A review of equipment and maintenance
practice is an enlightening part of any supplier audit.

Quote
We have had our current wave solder machine for about six years and it still looks new
and operates very well with almost no down-time. The secret is maintenance and
operator pride.
I used to do consulting on wave solder machines and often saw them stuffed in little
sheds and dirty corners because they are stinky and dirty. We have ours right in front and
make it an important part of customer tours.
Another important thing we do is having an annual maintenance done by the factory.
This gives a new set of eyes to look at the machine for little things that we may overlook
day to day. Also the technician gets to see all kinds of problems with poorly maintained
machines in the field and can help you dodge a bullet.
When the factory tech comes to perform maintenance on our machine they are always
amazed at the condition of our machine and tell us some horror stories of what they see
elsewhere.
Kenny Bloomquist on TechNet, 10 Mar 1999

There is no more to say!

Fluxer

Sealed spray fluxer systems need little more maintenance than checking the free
operation of any mechanical movements and keeping the feed tubes and
nozzles unblocked. Problems in those areas result in inconsistent flux coverage,
a fault that is soon apparent and easily diagnosed.
The same cannot be said of wave and foam fluxers, where the two mechanisms
that can affect the solder joint take place over time:
the density of the flux increases as solvents evaporate, producing a
more viscous flux and a thicker flux deposit
the flux gradually degrades through exposure to air, especially if the
environment is hot and/or humid.

The first of these is normally addressed by monitoring the density (also referred
to as ‘specific gravity’) of the flux. This relies on the difference in density
between the flux and the thinners used. The second problem area can only be
dealt with by regular (fortnightly or monthly) cleaning of the flux tank and
replacement of the flux.

Solder

Here there are three topics to consider:


having enough solder to ensure that the wave height is correct
making sure that the joints are not adversely affected by the build-up
of oxide on the solder surface
ensuring that the solder in the pot has not become contaminated.

The solder level in the bath must be maintained, and the pot replenished,
which means that the level of solder in the pot must be monitored regularly,
either automatically or by the operator.
Dross is formed when solder is moved or disturbed, and the oxide skin formed
by reaction with the oxygen in the atmosphere breaks up and mixes with the
solder underneath. Because the basic principle of wave soldering involves
moving solder around and letting it fall back into a bath of molten material, the
formation of dross is unavoidable unless an inert atmosphere is used. Allowed
to build up, dross has a number of negative effects, so it needs to be removed
regularly: depending on the application, some intervention may be needed every
hour!
For hand and reflow soldering, purity of the starting material is not an issue
since fresh solder is used. However, with wave soldering, solder in the bath is
continually reused, and may gradually pick up contamination dissolved from the
product being processed. Silver (from passive component terminations), copper
(from boards without a nickel barrier layer) and gold (from boards with nickel-
gold finishes) are the materials most frequently found. It is normal practice to
use a laboratory to carry out an analysis of the bath on a regular basis (perhaps
three-monthly) and then make the necessary additions of tin, to replace tin
oxide lost in the dross, or even replace the whole of the solder in the bath if it is
contaminated.

You might also like