You are on page 1of 19

Development of an empirical model to describe the local

high temperature corrosion risk of 13CrMo4-5 steel in


biomass CHP plants regarding the fuel wood chips

Dipl.-Ing. Thomas Gruber 1, Dr.-Ing. Kai Schulze 1, Dipl.-Ing. Dr. Robert Scharler 1,2,3, Dr. Barbara
Waldmann 4, Prof. Dr. Ferdinand Haider 5, Prof. Dipl.-Ing. Dr. Ingwald Obernberger 1,2,3
1
BIOENERGY 2020+ GmbH, Inffeldgasse 21b, A-8010 Graz, Austria
Tel.: +43/316/8739228; Fax: +43/316/8739202; E-mail: thomas.gruber@bioenergy2020.eu
2
BIOS BIOENERGIESYSTEME GmbH, Inffeldgasse 21b, A-8010 Graz, Austria
3
Institute for Process and Particle Engineering, Graz University of Technology, Austria
4
Corrmoran GmbH, Buchenstraße 13, D-86179 Augsburg, Germany
5
Institute of Experimental Physics, University Augsburg, Germany

ABSTRACT
An empirical model for the estimation of the high temperature corrosion in biomass
furnaces and boilers, using the fuel wood chips, has been developed at the Austrian
Bioenergy Competence Centre BIOENERGY 2020+. At the present state, the model
considers the influence of the flue gas temperature and velocity as well as the surface
temperature of the steel. The model is validated within the flue gas velocity range of 2
– 10 m/s, a flue gas temperature range of 625 – 880 °C and a surface temperature
between 450 – 550 °C.
The empirical function is a combination of an Arrhenius function which describes the
exponential dependence on the surface temperature and the flue gas temperature and a
linear dependence on the flue gas velocity. At the present state the model is able to
describe the corrosion potential for the fuel wood chips reasonably accurate. In next
future the model will be extended towards the high temperature corrosion behaviour
of the fuels wheat straw pellets and waste wood.
Keywords: biomass combustion, high temperature corrosion, CFD simulation, online
corrosion measurement, deposition layer

1. INTRODUCTION AND OBJECTIVES


Ash related problems have a strong influence on the operation of biomass combustion
plants for the production of heat and electric power, resulting in unscheduled outages
and reduced economic efficiencies. Among these problems, material corrosion of steel
surfaces in the hot furnace, of radiative boiler walls and of convective heat exchanger
tube bundles (e.g. superheaters) is of major importance especially when firing
biomass fuels with high contents of chlorine and alkali metals (waste wood, as well as
agricultural and herbaceous fuels like straw and Miscanthus) but also for conventional
wood fuels (wood chips, bark) with respect to increasing the steam parameters and
thus the efficiency of future biomass CHP plants.
While considerable research has been done regarding high temperature corrosion in
waste to energy plants (an overview is given by [1], [2] and [3]) far less research has
been performed in the field of high temperature corrosion in biomass combustion

-1-
plants ([4], [5] and [6]) especially for wood fuels with low chlorine and alkali
contents. For these fuels the quantities of the local influencing parameters like flue
gas temperatures, steel surface temperatures, flue gas concentrations of chlorine and
sulphur species as well as structure and composition of the inner deposit layers on the
steel surfaces are unknown in most cases. In addition, due to the high complexity of
the underlying processes, which are not fully understood yet, no simulation tools
which sufficiently describe corrosion in biomass fired boilers in dependence on the
influencing parameters are available. Nevertheless some authors (Davis [7], Linjevile
[8] and Warnecke [9]) already published results regarding the CFD-aided estimation
of corrosion rates in coal and waste combustion plants. The authors used empirical
correlations and basic assumptions in order to calculate the deposit build-up on the
walls as a basis for the estimation of the corrosion potential. Due to the simplifications
used for the description of the deposit build-up in their work, the models are not able
to sufficiently describe corrosion processes depending on relevant influencing
parameters and therefore are not suitable for an application in biomass combustion
plants in general.
This work is a first step towards the development of an empirical high temperature
corrosion model which shall then be linked to an already existing deposit formation
model [10] to estimate the local high temperature corrosion risk in biomass fired
furnaces and boilers. In this first step corrosion probes have been used to investigate
the corrosion rates for the common superheater steel 13CMo4-5 and the fuel wood
chips regarding the influencing parameters: flue gas temperature, velocity and steel
surface temperature. The next step will be an extension of the model towards two
other fuels, namely chemical untreated wheat straw pellets and waste wood. Note that,
even if corrosion rates are measured within the experimental test runs, the final
empirical model will only be able to give estimations regarding the local corrosion
potential.

2. Methodology

2.1 Overall overview of the test runs


The first campaign of test runs, using a corrosion probe (as described in detail in
chapter 2.2) to determine the most relevant influencing parameters for high
temperature corrosion in biomass furnaces, took place within a time period of 8 weeks
at the specially designed packed bed furnace of BIOENERGY 2020+. Dedicated
accompanying measurements have been performed in the first two weeks of the test
campaign. In addition balance test runs have been performed for the reference load
and the full load operation mode. The timetable of the test runs is shown in Table 1.
The discontinuous, accompanying measurements as well as the continuous corrosion
measurements are explained in the chapters 2.2 to 2.8.
Starting form week four parameter variations regarding the flue gas temperature and
velocity as well as the surface temperature have been performed. The flue gas
temperature was varied between 625 and 880 °C while the probe surface temperature
was varied between 450 and 550 °C to simulated nowadays and future life steam
temperatures of biomass CHP plants. The reference flue gas velocity was chosen to
2.8 m/s to reduce erosion processes and to ensure a sufficient deposit build-up within
the test period. To investigate the dependence on the flue gas velocity, variations from
3 to 10 m/s have been done.

-2-
Table 1: Time schedule of the test run
week of test run 1 2 3 4 5 6 7 8
start phase (corrosion layer build-up)
probe surface temp. variations
flue gas temp. variations (TFG = 850 °C)
flue gas temp. variations (TFG = 650 °C)
flue gas velocity and short term variations
balance test runs
dust measurements
aerosol measurements
deposition probe measurements
HCl/SOx measurements

The first three weeks were used to build up a stable corrosion layer which allows the
measurement of a reproducible corrosion signal. Therefore, starting with week four
the corrosion signal gained was used to develop the empirical function. Within week
four surface temperature variations have been performed. Starting at the pre-defined
reference state of 480 °C corrosion probe surface temperature (TS) and a flue gas
temperature at the corrosion probe (TFG) of around 750 °C, TS was varied to 450, 500,
520 and 550 °C. Each TS was investigated on a separate day. In week five TFG was
changed, starting from the reference state, to 850 °C, in week six TFG was changed to
650 °C. The flue gas temperature variation is achieved by a change of the electrical
heating power of the drop tube (see chapter 2.2) at unchanged load conditions of the
furnace. Within these two weeks the corrosion probe surface temperature was also
changed to achieve a widely spread set of data points. In test week seven the flue gas
velocity was changed from 2.8 m/s, at the reference condition, to seven respectively
10 m/s. To achieve this velocity increase on the one hand side the furnace load was
increased and on the other hand the mass flux was increased by a recirculation of the
flue gas. In addition to this, also some short term variations of all parameters have
been performed, which have not been taken into account for the development, but for
the validation of the empirical function. In test week eight balancing test runs at the
reference and the full load operation mode have been performed.

2.2 Online corrosion measurements


The aim of the test runs was the identification and quantification of the main
influencing parameters on the high temperature corrosion process using wood chips as
fuel and the steel 13CrMo4-5 as corrosion probe ring material, which is a commonly
used steel for superheaters in biomass CHP plants. Table 2 shows the chemical
composition of the steel whereas Table 5 shows the composition of the fuel used.

Table 2: Chemical composition of the steel 13CrMo4-5

C Si Mn P S N Cu Cr Mo Fe
min [weight-%] 0.08 0.40 0.70 0.40 remain
max [weight-%] 0.18 0.35 1.00 0.03 0.01 0.01 0.30 1.15 0.60 remain

The experiments have been carried out using a combined packed bed / drop tube
reactor, with which well-defined conditions regarding flue gas temperature,

-3-
composition and velocity as well as temperatures at the steel surface, typically
prevailing in biomass combustion plants and boilers can be achieved at continuous
operation mode. The used setup consists of a packed bed reactor (biomass grate
furnace equipped with air staging and flue gas recirculation), an upper transition part
to a successive heated vertical tube with isothermal conditions (the so-called drop
tube). The drop tube can be heated up (3x 20 kW electrical input power) to
temperatures of 1200 °C which allows a heating of the flue gas up to 900 °C at the
reference load operation mode of the reactor. To achieve flue gas temperatures of 625
°C the electrical heating of the drop tube is turned off. An additional water cooled
pipe is installed to bypass the measurement unit during start-up and shutdown of the
reactor. The drop tube has a length, which is sufficient to get a fully developed flow at
the entrance of the measurement port, which is equipped with a corrosion and/or a
deposition probe. At the position of the measurement port an additional optical port is
installed to enable a visual monitoring of the corrosion probe during the test runs.
Therefore, the reactor enables corrosion measurements under defined flow and
temperature conditions with a flue gas typical for fixed bed combustion systems.
Finally, the flue gas is transferred via a lower transitional part to a water cooled heat
exchanger before it enters the chimney. Before the entrance to the chimney an
additional port is installed to allow impactor-, dust and flue gas analysis
measurements.
water cooled flue
gas bypass
commentary:
TI: temperature measuring point
PI: temperature measuring point
FI: mass flow sensor
grate furnace
12-50 kW

secondary air
fuel
primary air
flue gas to electrical heated
chimney drop tube (3 x 20 kW)
flue gas recirculation

measuring point
for optical port for visual
impactor-, dust monitoring of the probe
and flue gas
analysis
measurements
measurement port for gas analysis,
dust-,deposition- and corrosion
water cooled heat measurements
exchanger: 50 kW

Figure 1: Schematic view of the combined packed bed / drop tube reactor with relevant units and
measurement ports
The corrosion probe provided by the company CORRMORAN GmbH (Augsburg,
Germany) is based on a system which has been developed at the Institute of Physics,
Augsburg University. The system is especially designed to simulate the high
temperature corrosion occurring on CHP superheater tubes. The corrosion probe is
inserted into the flue gas of the reactor at the measuring port below the heated drop
tube (see Figure 1).

-4-
The probe design consists of an air-cooled support lance made from a heat resisting
nickel based super-alloy (Figure 2 left), with an air cooled probe head (Figure 2 right).
The probe head itself is composed of four sample rings that are separated by ceramic
rings. Those sample rings, made of 13CrMo4-5, are held at a constant but due to the
air cooling arbitrary temperature (ring temperatures investigated during the test
campaign 450, 480, 500, 520 and 550 °C). When exposed to the flue gas an
electrolyte layer, which allows the measurement of the linear polarisation resistance,
is formed.
For the measurement a voltage is applied between the working electrode and the
counter electrode resulting in a current between the two electrodes depending on the
current corrosion rate. Since the chemical reactions on the electrodes affect the
measurement of the applied voltage, an additional reference ring, which is connected
to the working electrode, is necessary to measure the actual voltage between the
working and the counter electrode.
The applied voltages are shifted around 0 V, which results in different currents and
allow the recording of a current-voltage-characteristic. The linear content of the slope
of the characteristic is called the linear polarisation resistance
This resistance is directly proportional to the current corrosive attack. The measured
signal can be related to a corrosion rate, which describes the material loss over time
mm/1000h, subsequent to the test campaign by quantifying the loss of weight of the
ring M. A more detailed description of the probe can be found in [3].

(1) working electrode (3) counter electrode


(2) reference electrode (M) mass loss ring

ceramic rings

Figure 2: Schematic view of the corrosion probe (left) and the sensor probe head (right) with
three electrodes (1 - 3) and the mass loss ring (M) [11].

2.3 Flue gas analysis


A slip stream of the flue gas is sucked through a heated sampling probe. Included in
the sampling probe is a ceramic filter which separates the particles from the flue gas.
Then the flue gas passes over a heated pipe before it enters the analyser
(Rosemount/NGA 2000) which continuously measures the concentrations of O2, CO2
and CO in the flue gas.
Flue gas analyses have only been performed during the balance test runs.
2.4 Aerosol measurements
A Berner-type low-pressure impactor (BLPI), Type Hauke LPI30/0.0625/2, was used
to investigate the aerosol concentration in the flue gas. The measurements have been

-5-
performed after the heat exchanger (see Figure 1 for exact position). For measurement
purposes a slip stream of the flue gas is sucked isokinetically through the impactor,
which is preheated to 120 °C to avoid flue gas condensation. The impactor itself
consists of several stages. In each stage the flue gas changes its flow direction and
particles which are too big to follow the streamlines of the flue gas, impact on
aluminium sampling foils. These foils are weighted before and after the measurements
to determine the aerosol concentrations and are subsequently analysed by means of
wet chemical analysis. Further information is given in [12].
BLPI- measurements have been performed three times in the second week of the
corrosion test run at the measuring port located after the heat exchanger at the
chimney. The BLPI and all other accompanying measurements have exclusively be
performed under standard operating conditions (see chapter 3.1).
2.5 Total dust measurements
The total dust concentration in the flue gas was measured according to VDI 2066. The
used device is a Ströhlein ST E 4. For a certain period a slipstream of the flue gas is
sucked through a quartz wool filter, where the dust particles are precipitated. To
achieve representative sampling, the slip stream must be taken under isokinetic
conditions. For further information see [12].
Total dust measurements have been performed three times in the second week of the
corrosion test run at the measuring port located after the heat exchanger at the
chimney.
2.6 HCl/SOx measurements
The determination is carried out according to VDI 3480, Sheet 1. Flue gas is sucked
from an extraction point over several successively connected washing flasks with
distilled water and diluted sodium hydroxide solution (enriched with H2O2). The
dissolved anions are measured by ion-chromatography (IC).
HCl/SOx measurements have been performed five times in the first week of the test
run at the corrosion probe port (see Figure 1).
2.7 Deposit probe measurements
To gain more information about the chemical composition, the build-up rate and the
structure of the deposit layer, deposit probe measurements have been performed. The
deposit probe consists of a carrier lance with a test ring on top, which is cooled by air
and is inserted into the flue gas at the position of the corrosion probe. The air cooling
allows the investigation of arbitrary, constant ring temperatures, which simulate
conditions at superheaters with different life steam temperatures. A more detailed
description of the deposit probe measurement can be found in [13].
The investigation of different probe temperatures allows an estimation regarding the
condensation of different gaseous components such as KCl.
The mass gain of the test ring is determined by gravimetric measurements of the test
ring before and after the exposure to the flue gas. Scanning electron microscopy

-6-
(SEM) and energy dispersive X-ray (EDX) analysis have been performed to analyse
the deposits subsequently to the measurements.
Two deposit probe measurements have been performed, one at a probe temperature of
550 °C and the other one at the reference surface temperature of 480 °C. Both samples
were taken at standard operation conditions (see chapter 3.1) which mean a flue gas
temperature around 730 °C and a flue gas velocity of 2.8 m/s at the deposit probe. The
sampling time was in each case two hours.
2.8 SEM/EDX- analysis and sample preparation
To investigate the chemical composition of the deposit layer by means of SEM/EDX-
analysis, the deposit probe rings were first coated with a thin carbon layer to avoid an
electrical charging of the probe. The samples were analysed using a ZEISS Gemini
982 field emission scanning electron microscope equipped with a Noran Voyager X-
ray analysis system (Si(Li) detector, ultra-thin window). The oxygen content was
determined by stoichiometric calculations. The composition of the deposit layers were
determined by area scans using magnifications of 100 respectively 500 to compare the
overall compositions with local compositions. Because of the big interaction volume
of the electron beam with the sample not only the thin deposit layer becomes
analysed, but also the underlying steel of the probe ring. This is especially a problem
for thinner deposit layers at the leeward side of the probe ring, Therefore, the
elements Fe, Cr and Ni were excluded from the analyses since these are the main
components of the probe steel ring.
To gain more information regarding the corrosion mechanism occurring in biomass
combustion plants using the fuel wood chips, SEM/EDX–analyses of the corrosion
probe ring have been performed subsequently to the test runs. To investigate the
element distribution of the corrosion probe ring profile, first the ring was embedded in
epoxy resin, afterwards the profile was grinded with sand papers with decreasing
grain sizes until a smooth, plane surface was achieved. It was not possible to polish
the ring, since the organic solvent would attack the deposit layer of the corrosion
probe ring. Afterwards, the ring was also coated with a carbon layer. Element
mappings have been performed at a magnification of 400 to analyse the chemical
composition of the corrosion layer. The used systems for the mappings are a Zeiss
Ultra 55 field emission scanning electron microscope equipped with an EDAX
Pegasus X-ray analysis system. These element mappings show the distribution of
each element over the area investigated (bright colour means a high concentration,
weak colour a low concentration). Within these element mappings the oxygen content
was measured and not calculated stoichiometricaly to detect oxides and to achieve a
clear separation between the corrosion layer and the original steel surface. To increase
the accuracy of the EDX-analyses, the SEC- (standardless element coefficient) factors
of the EDX-analysis system were calibrated by reference to a Fe2O3 and a Fe3O4
standard sample.

2.9 Fuel and ash analyses


To investigate the fuel and ash composition representative samples of the grate ash
and of the fuel have been prepared. The moisture content of the fuel sample is

-7-
determined according to ÖNORM CEN/TS 14774. Fuel sample preparation is carried
out according to CEN/TS 14780. The ash content is determined according to CEN/TS
14775. Additionally, the corrected ash content has been calculated by reduction of the
ash content by the amounts of carbonates in the ash. This carbonates are formed
during the ashing process, which occurs at a constant temperature of 550 °C. In fixed
bed biomass furnaces carbonates do typically not occur due to the high temperatures
prevailing. The determination of C, H and N is carried out according to ÖNORM
CEN/TS 15104. The chlorine content is determined according to ÖNORM CEN/TS
15289. Major and minor ash forming elements are determined by multi-step
pressurised digestion of the fuel with HNO3 (65%) / HF (40%) / H3BO3 followed by
inductively coupled plasma optical emission spectroscopy (ICP-OES) or inductively
coupled plasma mass spectroscopy (ICP-MS) depending on detection limits.

3. Results
3.1 Combustion conditions
To determine the combustion conditions during the corrosion test runs, balancing test
runs have been performed at two different load conditions subsequent to the corrosion
test runs (see Table 1). Within these test runs, the fuel consumption, the amount of ash
produced as well as their chemical compositions was determined. Based on these
information mass and element balance calculations have been done and release rates
have been calculated to determine the combustion conditions and concentrations of K,
Cl and S in the flue gas at the corrosion probe. Moreover, additional measurements
have been executed, at the reference state conditions (23.4 kW operation mode), to
gain more information about the total dust and aerosol emission at the chimney as
well as the concentration of gaseous HCl and SOx at the corrosion probe. The
characteristic data of the different operation modes can be found in Table 5. To check
the results of the balance test runs and the temperature profile measured also a CFD
calculation of the biomass fired drop-tube reactor at standard operating conditions has
been performed. The results regarding the flue gas velocity and flue gas temperature
field near the corrosion probe are presented in Figure 3.

Figure 3: Calculated flue gas temperature field [°C] (left) and flue gas velocity field [m/s] (right)
at the corrosion probe (cross section through the axes of the reactor)

-8-
Table 3: Standard operating conditions of the furnace for the corrosion measurements at the
reference state
reference state combustion flue gas velocity variation state
condition combustion condition
parameter value unit value unit
fuel power input 23.4 kW 42.1 kW
fuel wood chips wood chips
fuel flow rate 6.3 kg/h 11.3 kg/h
temperature primary combustion zone 818 °C 804 °C
temperature secondary combustion zone 896 °C 940 °C
flue gas temperature after boiler 115 °C 190 °C
CO2 11.1 vol.-% FG d.b. 12.3 vol.-% FG d.b.
CO 13.8 mg/Nm³ 11.3 mg/Nm³
O2 9.5 vol.-% FG d.b. 8.4 vol.-% FG d.b.
primary air volume flux 15.4 Nm³/h 20.5 Nm³/h
secondary air volume flux 16.1 Nm³/h 21.8 Nm³/h
combustion air flux 31.5 Nm³/h 42.2 Nm³/h
flue gas mass flux 49.7 Nm³/h 83.5 Nm³/h
flue gas recirculation mass flux - Nm³/h 21.1 Nm³/h
air ratio primary combustion zone (with flue gas recirculation) 0.9 - 1.0 -
air ratio secondary combustion zone (with flue gas recirculation) 1.5 - 1.4 -
adiabatic flame temperature (with flue gas recirculation) 1245 °C 1083 °C

The results of the total dust, aerosol and the HCl/SOx measurements are presented in
Table 4. One can see that the main part of the dust emissions is formed by aerosols.
Table 5 shows the chemical composition of the fuel, the grate ash and the aerosols. In
each case one representative sample was analysed subsequently to the test runs.

Table 4: Flue gas emissions at standard operating conditions

emissions position mean std.-dev. unit


total dust chimney 24.3 5.8 mg/Nm³ d.b. 13%-O2
aerosols chimney 15.5 0.3 mg/Nm³ d.b. 13%-O2
HCl (gasous) corrosion measuring port 6.6 2.8 mg/Nm³ d.b. 13%-O2
SOx (gasous) corrosion measuring port 8.6 3.8 mg/Nm³ d.b. 13%-O2

Table 5: Concentrations of ash forming elements in the fuel wood chips, the grate ash and the
aerosols
element unit wood chips grate ash aerosols
TW w.t.% w.b. 21.40
AC w.t.-% d.b. 1.90
AC corr w.t.-% d.b. 1.41
C w.t.-% d.b. 48.10
H w.t.-% d.b. 5.87
N w.t.-% d.b. 0.26
TIC mg/kg d.b. 25,200 n.a.
TOC mg/kg d.b. 32,200 n.a.
S mg/kg d.b. 265.00 1,540 146,000
Cl mg/kg d.b. 105.00 94 66,900
Ca mg/kg d.b. 6,860.00 249,000 3,310
Si mg/kg d.b. 3,250.00 147,500 3,980
Mg mg/kg d.b. 1,040.00 23,150 618
Al mg/kg d.b. 585.00 20,550 n.a.
Na mg/kg d.b. 74.70 3,070 9,150
K mg/kg d.b. 1,780.00 67,000 379,000
Fe mg/kg d.b. 328.00 15,400 n.a.
P mg/kg d.b. 189.00 10,250 2,440
Mn mg/kg d.b. 63.20 6,340 221
Zn mg/kg d.b. 19.40 130 21,000
Pb mg/kg d.b. 1.50 < 30 n.a.

The release rates for K, S and Cl could be determined to 18 % for K, 90 % for S and
98 % for Cl. They have been calculated from the difference between the mass of these
elements found in the grate ash and the amount of these elements supplied with the
fuel. The concentrations at the position of the corrosion probe can be found in Table
6. Even though the gaseous concentration of Cl is slightly higher than the maximum
amount of Cl estimated by the release calculation, the error lies within the standard

-9-
deviation of the HCl measurement (see Table 4) and clearly confirms that almost all
Cl in the fuel evaporates. The amount of K in gaseous form at the probe cannot be
defined because no high temperature impactor measurements have been performed.

Table 6: Concentration of K, S and Cl at the position of the corrosion probe. The released
amount is based on the release rates determined, the amount of gaseous Cl and S is based on the
HCl/SOx measurements

released amount gaseous amount


3 3
mg/Nm d.b. 13%O2 mg/Nm d.b. 13%O2
Cl 5.7 6.4
S 20.6 4.3
K 21.6 not determined

3.1 Corrosion probe measurements


It took approximately three weeks to build up a stable deposition and corrosion layer
(see Figure 4), which is necessary to achieve a stable, reproducible corrosion signal.
Horn [14] reported that it takes around 10 days to achieve a stable layer structure in
similar measurements, performed in waste to energy CHP plants. Therefore only the
last 20 days of the test series have been taken into account for the development and
testing of the empirical high temperature corrosion model. The test series lasted 1070
h in which the biomass furnace was 325 h in operation. Due to safety reasons the
furnace had to be shut down in the evenings and at weekends. To prevent surface
erosion of the corrosion layer by thermal stress the drop tube was heated overnight at
a constant temperature of 850 °C at the heating elements, which led to a constant
temperature of about 540°C at the measurement port.

after 154 h after 346 h after 634 h after 850 h

Figure 4: The pictures show (from left to right) the development of the deposition layer during
the test series on the corrosion probe. The pictures were taken after 154, 346, 634 and 850 hours
of measurement, the total duration of the test run was 1070 hours.
To study long-term changes of the corrosion signal, all parameter variations started
from a pre-defined reference state (flue gas temperature around 730 °C, probe surface
temperature of 480 °C, flue gas velocity of 2.8 m/s). This reference state was re-
established at the beginning and the end of each day. At the end of a test day the
reference corrosion signal is typically slightly higher than in the morning (see Figure
5; left). This effect can be explained by a slow increase of the surface temperature of
the surrounding walls of the measurement unit which influences the corrosion probe
as well as the temperature sensors (thermocouples) by radiation slightly.
The measurements clearly show a correlation of all parameters investigated (flue gas
velocity, flue gas temperature and probe surface temperature) with the high
temperature corrosion rate (see Figure 5).

-10-
Moreover, the corrosion signal clearly shows that the corrosion rate strongly
decreases immediately after shutdown of furnace and further decreases to zero
overnight. These trends could be reproduced over the various following days.
The total mass loss of the ring M (see Figure 2) during the field campaign could be
determined to 0.44 g by gravimetric measurement of the test-ring before and after the
exposure to the flue gas. This corresponds to a total material loss of 0.04 mm. The
thickness distribution of the corrosion layer is discussed in detail in chapter 3.6.

vFG = 2.8 m/s vFG = 7.0 m/s vFG = 2.8 m/s

reference state

Figure 5: Example of a measured corrosion rate over a testing day with the calculated corrosion
rate and the relevant temperatures (probe surface and flue gas temperature) used in the
empirical model; in the upper figure the flue gas velocity vFG varied between 2.8 to 7.0 m/s in the
lower figure the flue gas velocity vFG was kept constant at 2.8 m/s. The lower figure represents a
day which was used to validate the empirical function, therefore no reference state is marked.
The calculated corrosion rate is based on the developed empirical function presented in chapter
3.5.

-11-
3.2 Deposit probe measurements
SEM/EDX-analyses have been performed on three different locations of the corrosion
ring: in the luv, luv+50 degrees and the leeward direction of the flue gas. The results
of the SEM/EDX-analyses are presented in Figure 6. It can be shown that the
deposition layer mainly consists of potassium and sulphate which indicates the
presents of potassiumsulphate based on the molar ratios measured. Based on the
analyses chlorine can only be found in very small amounts in case of 480 °C surface
temperature, with a slightly higher concentration on the leeward-side of the probe
ring. The build-up rates measured for the deposition layers are: 2.81 g/m²h in case of
480 °C surface temperature and 1.39 g/m²h for 550 °C surface temperature. These
values can be compared with similar measurements performed in a medium scale (440
kWth nominal capacity) biomass grate combustion plant [15] using wood chips as
fuel. This work reports total dust emissions between 180-380 mg/Nm³ and aerosol
emissions between 13-17 mg/Nm³ related to dry flue gas and 13 Vol.-% O2. The total
dust emissions are considerably higher than those measured within this work.
Therefore, also the build-up rates presented in [15] are with 9.6 g/m²h in case of 550
°C and 7.1 g/m²h in case of 450 °C surface temperature (both measurement lasted 2
hours) higher than those found within this study. The results clearly show that in
larger grate combustion plants the concentration of coarse fly ash emissions increases
which also influences the deposit formation behaviour (see also [13]).
O
100% Pb
Zn
80%
[Atomic-%]

Ti
60% S
40% Cl
P
20% K
500 x
0% Na
100x 500x 100x 500x 100x 500x Al
Si
luv luv+50 lee
Ca
surface temperature = 480°C (duration 2h) Mg

O
100% Pb
80% Zn
[Atomic-%]

Ti
60%
S
40% Cl
20% P
0%
K 500 x
Na
100x 500x 100x 500x 100x 500x
Al
luv luv+50 lee Si
surface temperature = 550°C (duration 2h) Ca
Mg

Figure 6: The upper left part shows the results of the EDX- analyses for a probe surface
temperature of 480 °C; the lower part shows the result for a probe surface temperature of 550
°C. To visualize the deposit layer structure, the right side shows overview SEM-pictures, at a
magnification of 100, of the luv-side.

3.4 EDX-analysis of the corrosion probe ring


Figure 7 clearly shows that the deposit layer mainly consists of potassium, oxygen
and sulphur, which indicates the compound potassium sulphate, while the corrosion
layer mainly consists of iron oxides with an increased sulphur concentration at the
corrosion front (metal-oxide interface). Chlorine could only be found in small

-12-
amounts in deposit probes at lower temperatures (see chapter 3.2) and on a single spot
on the lee-side of the corrosion ring. Since many authors (for example [16] and [17])
report a compact inner corrosion layer, consisting of FeCl2, it is assumed that another
corrosion mechanism instead of active Cl-induced oxidation seems to be of relevance
for the case investigated. Although the enrichment of sulphur at the corrosion front is
an indicator for a sulphur induced corrosion mechanism, further work has do be done
to identify the mechanism in more detail.
SEM Fe O K S Cl

deposit layer
luv
corrosion layer

steel

deposit layer
luv+50
corrosion layer

steel

deposit layer
lee corrosion layer

steel

Figure 7: SEM/EDX element mapping of the most important elements present on the corrosion
probe ring. The brighter an element appears the higher is the concentration of the element. Still
it is not possible to compare the quantities of different elements. Because of the lack of chlorine
one can only see the background noise of the EDX analyses in the Cl mappings. The only
exception is the small chlorine containing area on the leeward side of the ring.

3.5 Model development of an empirical corrosion potential


model
Based on the experimental data, an empirical function has been derived to describe
the corrosion rate as a function of the parameters: tube surface temperature, flue gas
temperature and flue gas velocity. The corrosion rates determined (see chapter 2.2)
show an exponential dependence on the flue gas temperature as well as on the surface
temperature.
Figure 8 shows on the left side the dependence on the flue gas temperature and on the
flue gas velocity (presented in different colours) for a constant surface temperature of
480 °C. On the right side one can see the exponential increase with increasing probe
surface temperatures TS.
Therefore as a first, mathematically motivated approach an Arrhenius-function was
chosen to describe the dependence on those parameters, which is a common approach
for reaction limited high temperature corrosion processes [3]. This approach is
approved by the good agreement between measured and estimated corrosion rates, as
can be seen in Figure 5 and the fact that the underlying corrosion mechanism is
unknown in the present case up to know. Moreover, it follows the general approach
proposed by Maisch [3] for waste incineration.

-13-
Figure 8: (Left) Empirical function for the two limiting flue gas velocities 2 and 10 m/s at a
constant surface temperature of 480 °C. (Right) Plot of the empirical function for various surface
temperatures and a flue gas velocity of 2.8 m/s in comparison to measurement values

Figure 9: Correlation between the parameters A(TS) and B(TS) and the surface temperature
after linearization.
The parameters of the empirical corrosion model have been determined by the
following stepwise approach:
 As a first step, the parameters A(TS) and B(TS) where optimised for different,
constant surface temperatures TS, namely 450, 480, 500, 520, 550 °C at a
constant flue gas velocity of 2.8 ± 0.5 m/s. Unfortunately there was an
insufficient amount of experimental data for probe temperatures of 450 and 500
°C, to develop a proper function.
 As a next step, the dependence of these parameters on the surface temperature
was investigated. It can be shown, that both parameters A(TS) and B(TS)
increase the corrosion rate exponentially. The correlation between the
parameters and the surface temperature after linearization can be found in Figure
9.
 The dependence on the flue gas velocity is taken into account by a linear
correction to the already- with respect to the other two parameters- optimised
empirical function. Some literature claims that the influence of the flue gas
velocity might also depend on the flue gas and steel surface temperature [3].
Within this work it was not possible to confirm this theory. Moreover, no further
improvements on the empirical function could be achieved by including a more

-14-
complex, than a linear dependency on the flue gas velocity. This assumption is
not based on any thermodynamic reason, but at this stage of development the
linear dependence on the flue gas velocity leads, from a mathematical point of
view, to the best empirical description on the basis of the smallest mean squared
error. A possible reason for this could be the small amount of data points used
for optimising the dependence on the flue gas velocity, or the already existing
uncertainties with respect to the flue gas and probe surface temperatures allow,
in this case, no more precise description. Therefore, the dependence on the flue
gas velocity was determined at a fixed surface temperature of 480 °C and a flue
gas temperature between 750 and 800 °C. Further work has to be done to
develop a more mechanistic approach regarding the velocity dependence.
Table 7: Basic empirical function to describe the high temperature corrosion rate in biomass
furnaces for the fuel wood chips
with:
 B TS   CorrRate = corrosion rate [mm/1000h]
CorrRate  ATS   exp    C v FG 
 R  TFG
A(TS) = max. corrosion rate [mm/1000h]

A(TS )  exp a  TS  b 
B(TS) = activation energy [J/mol]
TFG = flue gas temperature [K]
B(TS )  c  TS  d TS = surface temperature [K]
vFG = flue gas velocity [m/s]
C v FG   e  v FG  f
R = gas constant [J/(mol K)]
a-f = constants

The empirical model was validated by the corrosion probe experiments. Figure 5
shows a comparison between the measured and the calculated corrosion rate. The
results show that already at the present state of development the measured corrosion
rates are well reproduced. Even though the experimental data for the surface
temperatures 450 and 500 °C could not be used for optimisation purposes, they are of
major importance for the validation of the empirical function. As can be seen in
Figure 8 the empirical function is able to estimate the measured corrosion rates quite
well for these surface temperatures too.
Retschitzegger [13] reports corrosion rates, at reference state conditions, that are
about 50 % lower than those found within this work. A possible reason for that could
be a thicker deposit and corrosion layer (see chapter 3.6), which might decrease the
material transport of corrosive species to the corrosion front as these measurements
have been performed at a wood chip fired grate equipped with a water tube steam
boiler (nominal fuel power input 28 MW) where coarse fly ash emissions are much
higher.
3.6 Thickness measurements of the corrosion layer
Since the corrosion layer, the deposit layer and the steel have different structures and
chemical compositions, the thickness of the different layers can be determined by
SEM-analysis. Figure 10 shows an example of such SEM-analysis. The thickness of
the corrosion layer was measured at seven positions located on one half of the probe
ring. The other half of the ring has, due to symmetry reasons, the same layer structure.
At each position the thickness of the corrosion layer was determined at least two

-15-
times. The distribution of the average thickness at each ring position is presented in
Figure 10 on the right hand side.

90

80

70

60

Thickness [µm]
50

40

30

20

10

0
0 20 40 60 80 100 120 140 160 180
Tilt angle from the luv-position [°]

Figure 10: (Left) Exemplary picture of a layer thickness measurement. The red lines show the
thickness of the corrosion layer, the blue lines of the ash deposit layer. (Right) Distribution of the
corrosion layer thickness over one side of the probe ring: An angle of 0° represents the luv-
side;and of 180° the lee-side
It can be seen that there is no clear dependence of the thickness of the corrosion layer
on the ring position. This results in an average corrosion layer thickness of 56 ± 20
µm. The measured value can be compared with the estimated thickness of the
corrosion layer according to the measured mass loss of the ring M (Figure 2), which
implies a corrosion layer thickness of 87 µm (assuming a corrosion layer consisting of
Fe2O3 with a density of 5.25 g/cm³ [18]). The deviation can be explained by the high
variation of the SEM- thickness measurements and the assumption that the other half
of the ring has the same structure due to symmetry reasons.
The structure of the deposit layer is highly inhomogeneous as can be seen in Figure
10. The mean deposit layer thickness could be determined to 41 ± 26 µm, which is
rather thin compared to the results Retschitzegger [13] has found.

4. SUMMARY AND CONCLUSIONS


As first step towards a CFD- based empirical high temperature corrosion model online
corrosion measurements have been performed at a specially designed packed bed/
drop tube reactor using the fuel wood chips. These measurements show a clear
dependence on the parameters flue gas temperature and velocity as well as steel
surface temperature. Based on the experimental data gained an empirical function has
been developed which is a combination of an Arrhenius-function (which describes the
dependence on the flue gas and the steel surface temperature) and a linear dependence
on the flue gas velocity. This empirical function is able to reproduce the corrosion
rates measured quite well. The function has been validated for a flue gas velocity
between 2 and 10 m/s, a flue gas temperature range of 625 – 880 °C and steel surface
temperatures between 450 – 550 °C.
Moreover, accompanying measurements have been performed to investigate the
corrosion mechanism prevailing more carefully. Deposition measurements as well as
subsequent SEM/EDX investigations of the corrosion probe rings show:

-16-
 A lack of chlorine at the corrosion front
 A deposition layer mainly consisting of potassium sulphate
 A small amount of chlorine in the deposit sample taken at 480 °C
surface temperature. No chlorine could be found in the deposit sample
taken at 550 °C. The lack of chlorine at the corrosion probe rings may
be due to the sulphation of chlorides or to the temperature changes of
the ring surface temperature.
 An enrichment of sulphur at the corrosion front is obvious
Therefore, it is suggested that the leading corrosion mechanism is different from the
well known active Cl-induced oxidation caused by an inner FeCl2 layer at the
corrosion front. Further work has to be done to identify the prevailing corrosion
mechanism for chemically untreated wood chips as fuel. It seems that a sulphur
induced corrosion dominates which is also confirmed by comparable measurements of
at large-scale fixed bed systems fired with wood chips [13].
As a plausibility check for the corrosion rates determined, thickness measurements
have been performed, using SEM-analysis. Both corrosion layer thicknesses
determined by the mass loss as well as by SEM-analysis are in good agreement.
The corrosion rates measured within this work can be compared to corrosion
measurements performed in large-scale biomass grate combustion plants [13]. The
rates determined differ by a factor of about 0.5 for comparable fuel qualities. A
possible explanation why the corrosion rates determined at large-scale plants are
lower may be the large difference of the deposit and corrosion layer thickness due to
the considerably larger concentration of coarse fly ash particles in the flue gas.
For the boiler steel investigated and the fuel wood chips (spruce), an empirical
corrosion model is now available for an estimation of the local high temperature
corrosion potential in biomass CHP plants.

5. Outlook
The next step in the development process towards a basic empirical corrosion model
is the evolution of the influence of different ratios of S to Cl by using different
biomass fuels. Therefore, measurements are presently ongoing using the fuels
chemically untreated wheat straw pellets and quality sorted waste wood (according to
the German waste wood categories).
Subsequently, the development of a more sophisticated CFD-based corrosion model is
planed, which considers transport processes and chemical reactions between the steel
surface, the surrounding deposit layer and the gas phase for the most relevant high
temperature corrosion processes determined for biomass grate combustion plants,
which shall be achieved by implementing a corrosion model in an already developed
ash deposit formation model [10].

-17-
6. ACKNOWLEDGEMENTS
The work presented was performed within the iK program of the BIOENERGY
2020+ GmbH. The financial support of the Austrian Research Promotion Agency
(FFG) (program: ModSim: project number: 828703) as well as of the companies BIOS
BIOENERGIESYSTEME GmbH, KWB Kraft und Wärme aus Biomasse GesmbH,
POLITECHNIK Luft- und Feuerungstechnik GmbH and Josef Bertsch GmbH & Co.
is gratefully acknowledged. The online corrosion probe measurements were
performed in cooperation with CORRMORAN GmbH, Germany. The good
cooperation with CORRMORAN GmbH concerning measurement support and
evaluation input is highlighted.

7. REFERENCES
[1] C. Schroer and J. Konys. Rauchgasseitige Hochtemperatur-Korrosion in
Müllverbrennungsanlagen - Ergebnisse und Bewertung einer Literaturrecherche,
Wissenschaftliche Berichte FZKA 6695, Forschungszentrum Karlsruhe GmbH, 2002

[2] B. Waldmann. Korrosion in Anlagen zur thermischen Abfallverwertung:


elektrochemische Korrosionserfassung und Modellbildung, PhD – Thesis, University
of Augsburg, 2007

[3] S. Maisch. Identifikation und Quantifizierung von Korrosionsrelevanten


Parametern in Müllverbrennungsanlagen mittels Charakterisierung der deponierten
Partikel und elektrochemischer Online-Messungen, PhD – Thesis, University of
Augsburg, 2011

[4] I. Obernberger, T. Brunner. Depositionen und Korrosion in Biomassefeuerungen,


Tagungsband zum VDI-Seminar 430504 Beläge und Korrosion in
Großfeuerungsanlagen, Göttingen, Deutschland, VDI-Wissensforum GmbH (Hrsg.),
Düsseldorf, Deutschland, 2004

[5] M. Ottmann. Verbrennung biogener Brennstoffe in stationären Wirbelschicht-


feuerungen, PhD – Thesis, TU – München, 2006

[6] F. Frandsen. Utilizing biomass and waste for power production - a decade of
contributing to the understanding, interpretation and analysis of deposits and
corrosion products, Fuel 84:1277-1294, 2005

[7] K. A. Davis, T. M. Linjewile, J. Valentine, D, Swensen, D. Shino. A Multi-point


Corrosion Monitorig System Applied in a 1300 MW Coal-fired Boiler, Anti-
Corrosion Methods and Materials 51:321-330, 2004

[8] T. M. Linjewile, J. Valentine, K.A. Davis, N. S. Harding. Prediction and real-time


monitoring techniques for corrosion characterisation in furnaces, Materials at high
temperatures 20(1/2), 2003

[9] R. Warnecke, S. Horn, M. Weghaus. Feuerungssimulation zur Aufdeckung


korrosiver Chloride, In: Tagungsband zur VDI-Veranstaltung 06KO005008, Beläge

-18-
und Korrosion, Verfahrenstechnik und Konstruktion in Großfeuerungsanlagen,
Oberhausen, Deutschland, VDI-Wissensforum GmbH (Hrsg.), Düsseldorf,
Deutschland, 2008

[10] K. Schulze, R. Scharler, M. Telian, I. Obernberger. Advanced modelling of


deposit formation in biomass furnaces – investigation of mechanism and comparison
with deposit measurements in a small scale pellet boiler, In: Proceedings of the
international Conference Impacts on Fuel Quality on Power production and
Environment, Lapland, Finland, EPRI (Ed.), Palo Alto,CA; USA, 2011

[11] C. Deuerling, B. Waldmann. Innovative Messtechnik zur kombinierten


Korrosions- und Materialforschung im Kraftwerksbereich, 43. Kraftwerkstechnisches
Kolloquium 2011, Dresden, 2011

[12] T. Brunner. Aerosols and coarse fly ashes in fixed-bed biomass combustion, PhD
– Thesis, TU – Graz, ISBN 3-9501980-3-2, 2006

[13] S. Retschitzegger, T. Brunner, I. Obernberger, B. Waldmann. Assessment of


online corrosion measurements in combination with fuel analyses, aerosol and deposit
measurements in a biomass CHP plant, In: Proceedings of the international
Conference Impacts on Fuel Quality on Power production and Environment,
Puchberg, Austria, to be published, 2012

[14] S. Horn, F. Haider, B. Waldmann, R. Warnecke. Quantifizierte Betrachtung des


Stoffübergangs in Belägen und Korrosionsschichten zur Beschreibung der
Korrosionsgeschwindigkeit, In: Tagungsband zur VDI-Veranstaltung 06KO005008,
Beläge und Korrosion, Verfahrenstechnik und Konstruktion in
Großfeuerungsanlagen, Oberhausen, Deutschland, VDI-Wissensforum GmbH (Hrsg.),
Düsseldorf, Deutschland, 2008

[15] MAWERA Feuerungsanlage GmbH.; Project no. SES-CT-2033-502679,


BIOASH, Ash and aerosol related problems in biomass combustion and co-firing,
Deliverable D10, Evaluation of test runs at the pilot-scale combustion unit at
MAWERA, 2006

[16] H. Reichel, U. Schirmer. Waste incineration plants in the FRG: Construction,


materials, investigation on cases of corrosion, Materials and Corrosion 40:135-141,
1989

[17] H. Nielsen, F. Frandsen, K. Dam-Johansen, L. Baxter. The implications of


chlorine-associated corrosion on the operation of biomass-fired boilers, Progress in
energy and combustion science 26:283-298, 2000

[18] CRC Press LLC, 2007: CRC Handbook of Chemistry & Physics, Version 2007

-19-

You might also like