Glycerine Purification
Glycerine Purification
Glycerine Purification
Authorisation A. Hoogendoorn
Dit project is uitgevoerd met subsidie van het Ministerie voor Economische Zaken; Besluit Energie Onderzoek Subsidie: Lange
Termijn (NEO)
This project was executed with a grant from the Dutch Ministry of Economic Affairs; Besluit Energie Onderzoek Subsidie:
Lange Termijn (NEO)
Ingenia © 2007
No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or any means, electronic, mechanical, photocopying,
recording, scanning or otherwise, except as with the written permission of Ingenia. This publication has been composed to provide accurate and
authoritative information in regard to the subject matter. However Ingenia is not liable for any direct, indirect, incidental or consequential damage, caused by
the use or application of the information of or data from this publication, or the impossibility to use or apply this information and/or these data. Ingenia is a
legally protected and registered trademark of Ingenia (Bureau Benelux des Marques dep.nr. 100.09.58) .
1 Introduction.....................................................................................................7
1.1 Objective....................................................................................................................7
1.2 Background................................................................................................................7
6.4.1 Hydrogen and Ethanol Production from Bacteria Enterobacter aerogenes HU-10160
7 Conclusions .................................................................................................. 81
Appendices
APPENDIX A Literature cited ................................................................................................................ 85
APPENDIX B Some glycerine market data by ADM Connemann (2003)............................................ 90
Figures
Figure 2-1 Schematic of commonly used glycerin splitting at biodiesel factories ................................... 9
Figure 2-2 Process chart of a continuous biodiesel plant by Energea.................................................. 10
Figure 2-3 Biodiesel production capacity in Germany 1998 – 2006 [UFOP.de] ................................... 12
Figure 2-4 Development of biodiesel production capacity and estimated glycerin production ............. 12
Figure 2-5 Reaction schematic of transesterification of triglycerides to biodiesel [4] ........................... 13
Figure 2-6 Conventional and biodiesel glycerin pathways and applications........................................ 15
Figure 2-7 End uses of glycerin with regional variations according to [7] ............................................. 16
Figure 2-8 Traditional glycerin applications [9]...................................................................................... 16
Figure 2-9 Bioking 200 kWth glycerin/bio-oil burner and boiler (left), gas turbine duct burner running on
crude glycerin (Heat Power & Ingenia; right) ................................................................................. 19
Figure 2-10 Impact of biodiesel glycerin on the glycerin market prices (99,5%, $/pound) ................... 21
Figure 3-1 Advantages & process characteristics of enzymatic biodiesel production according to
Lanxess [58]................................................................................................................................... 24
Tables
Table 2-1 Typical composition of crude glycerin from biodiesel production [8] .................................... 14
Table 5-1 Economic comparison of conventional versus enzymatic biodiesel production ................... 52
1.1 Objective
The goal of this survey is the execution of a feasibility study for the more cost effective upgrade of
glycerin and consequent transformation to high-quality products. Herein both the filter materials and
the side products like oil, biodiesel and soapstocks that arise during purification should be regenerated
(find a useful application). The original ambitions of the project partners were to come to:
• A crude glycerin that is inherently far less contaminated because of the addition of bio-
catalysts during the biodiesel production process;
• The feasible application of regenerative column adsorption techniques for further purification;
• A glycerin refinement that is twice as cost effective, more energy efficient and more simple;
• Regeneration of both adsorbentia and side products (oil, biodiesel and soap stocks);
• Identification of the most promising transformation routes to high-quality products (in chemistry
and pharmacy);
• The determination of the feasibility of a completely new industrial process with a high-quality
chemical or pharmaceutical end product.
1.2 Background
The regular production of biodiesel from oils and fats implies the production of about 12-15 wt% crude
glycerin as a side product. The EU production of biodiesel currently amounts to about 7 million tonnes
per year (production capacity of 11,5 Mtonnes; [FO Licht 2007]). It is expected that this amount will
double during the next five years, in line with the EU’s goals on energy security and sustainable
mobility (EG 2003/30 etc.).
The current annual amount of glycerin arising from this biodiesel production amounts to some 1,9
Mtonnes and will continue to rise proportionally. The existing world market for pure glycerol for high-
quality industrial applications (chemical and pharmaceutical) only covers some 0,9 -1,0 Mtonnes per
Crude glycerin (purity 50% - 90%), as it is produced during biodiesel production, unfortunately
contains too many contaminants to find a useful application in chemistry or pharmacy without
treatment. For example, the ash content can amount to several percents whereas only ppm levels of
contaminants may be allowed. As a result of the high purification cost of glycerin the application of
glcyerine in high-quality pharmaceutical and chemical applications is still only limited. The glycerin is
increasingly used in crude form, for instance as a fuel in cement kilns or exported to China to be used
there as a fuel in coal fired power plants, waste combustion installations and cement kilns.
It can be concluded that the current problems around glycerin are bivalent (and with increasing
severity):
1. Refinement of glycerin currently is very expensive and complex, which disqualifies the product
for high-quality use. The market is overwhelmed with crude glycerin with a price window of 0
to +150 Euro/tonne, with a low-quality application in combustion, digestion and a sharply
increasing export to USA and China.
2. The amount of glycerin production from biodiesel is at 1,9 Mtonnes/yr so high, compared to
the current market of 0,9 Mtonnes/yr, that additional high-quality chemical and pharmaceutical
applications need to be identified.
The European Union also recognises this issue and has reserved a separate amount of money for
research into useful applications of glycerin within the Seventh Framework Programme (FP7).
A typical biodiesel plant uses around 1-2 wt-% of KOH (wt-% on oil basis) as a homogeneous catalyst
while this catalyst remains as an unwanted pollutant in the crude glycerin. Most biodiesel plants are
very sensitive towards Free Fatty Acids, Phosphorous, polymers and water in the feed stock flows.
Figure 2.1 shows the glycerin upgrading from 50-60% purity towards 75-85% purity by means of soap
splitting. These kind of sophisticated process options are only used at bigger sized biodiesel plants
(typically > 50 ktonnes/yr). Figure 2.2 shows a rather sophisticated process flow chart of a biodiesel
plant built by the Austrian company Energea in which acid esterification and a splitting of the crude
glycerin into 80% glycerin, K2SO4 and Free Fatty Acids takes place. The Free Fatty Acids are used to
produce biodiesel in an acid catalysed esterification step (not so common).
FFA’s for
biodiesel
K2SO4
Fertiliser
World biodiesel production capacity will grow to around 23 Mtonnes (December 2007) while actual
production will probably be around 2/3 of that figure. These production numbers equate to around 1,9
Mtonnes of crude glycerin production at year end 2007 (figure 2.2).
35 3,00
production (Mtonnes/yr)
30
capacity (Mtonnes/yr)
Biodiesel production
2,50
Est. crude glycerine
25
2,00
20
1,50
15
1,00
10
5 0,50
0 0,00
Dec. 2006 Dec. 2007 Dec. 2008
Figure 2-4 Development of biodiesel production capacity and estimated glycerin production
In the table below some typical composition data for biodiesel-derived glycerin are given. Most of the
contaminants can be traced back to the biodiesel synthesis process, for example the unreacted
methanol that was not completely evaporated. Furthermore the concentrations of Na and K can tell
whether caustic soda (sodium hydroxide, NaOH) or potash lye (potassium hydroxide, KOH) was used
as a catalyst for the transesterification. Alkali metals like Na, K, Ca and Mg are naturally present in
vegetable oils. Sulphate and phosphate may remain from neutralisation of the mixture with sulphuric
or phosphoric acid.
Table 2-1 Typical composition of crude glycerin from biodiesel production [8]
The current annual global glycerin market is mainly dominated by high-purity and hence high-value
applications and amounts to 0,9 – 1,0 Mtonnes/yr. This glycerin is produced from both palm oil and
tallow by companies like P&G, Cognis, Uniqema, Vitusa, Dow Chemical and Dial. The glycerin
production from biodiesel exceeds the current conventional glycerin production by around 1
Mtonnes/yr (see figure 2.2) and, although of a lesser quality, is competing with conventional glycerin
production (see figure below).
According to [7] important applications are regionally determined. For Europe the largest discerned
single application is personal and oral care (22%). In the figure below the regional difference between
the size of this share in Europe, USA and Japan is well visible. In Japan the largest discerned single
Figure 2-7 End uses of glycerin with regional variations according to [7]
According to a presentation by Mr. Van Loo from the Dutch company Procede the traditional
applications of glycerin can be discerned as follows:
Drugs
• Used in medical and pharmaceutical preparations, mainly as a means of improving smoothness,
providing lubrication and as a humectants. Also may be used to lower intracranial and intraocular
pressures.
• Laxative suppositories and cough syrups.
Personal care:
• Used in toothpaste, mouthwashes, skin care products, hair care products and soaps.
• Serves as an emollient, humectants, solvent and lubricant in personal care products.
• Competes with sorbitol although glycerin has better taste and higher solubility.
• A component of glycerin soap.
Polyether polyols
• One of the major raw materials for the manufacture of polyols for flexible foams
and to a lesser extent rigid polyurethane foams.
• Glycerin is the initiator to which propylene oxide/ethylene oxide is added.
High-explosives
[11]
Nitroglycerin is extremely powerful. A mere 10 ml will expand 10,000 times into 100 litres of gas at
an explosive velocity of 7,700 metres per second (17,224 miles per hour) -- more powerful than TNT.
Love potion
Nitroglycerin's action as an effective vasodilator led in 1998 to the release of RESTORE, the first ever
fully tested, effective topical cream for the safe treatment of male erectile dysfunction (impotence).
Restore" contains 1% nitroglycerin and is "effective within minutes of application of achieving an
erection of up to 45 minutes duration.
Safe sweetener
Glycerin is an alcohol (glycerol) and is used as a preservative in the food industry, as well as a
sweetener: it is very sweet, yet it contains no sugar. This makes it an ideal sweetener for patients who
cannot take sugar, such as the increasing number of Candida sufferers. Vegetable glycerin is said to
be the "only acceptable sweetener" for Candida patients.
Health supplement
Health supplement for sportsmen -- Glycerin increases blood volume, enhances temperature
regulation and improves exercise performance in the heat, or so it is claimed. It helps "hyperhydrate"
the body by increasing blood volume levels and helping to delay dehydration. Following glycerol
consumption, heart rate and body core temperature are lower during exercise in the heat, suggesting
an ergogenic (performance enhancing) effect. In long duration activities, a larger supply of stored
water may lead to a delay in dehydration and exhaustion.
Burning glycerin
The glycerin by-product burns well, but unless it's properly combusted at high temperatures it will
release toxic acrolein fumes, which mainly form at between 200 and 300 º C (392-572 º F). At natural
gas prices around 9 €/GJ, crude glycerin will have a natural gas substitution value around 100-140
€/ton. The disadvantage of using crude glycerin as a fuel is that high dust emissions need to be
prevented and thus dust filters need to be used.
Figure 2-9 Bioking 200 kWth glycerin/bio-oil burner and boiler (left), gas turbine duct burner running on
crude glycerin (Heat Power & Ingenia; right)
Figure 2.7 shows the impact of biodiesel derived glycerin on the existing market. Both the year 2006
and 2007 are characterised by low glycerin prices. Current European prices for high-purity glycerin
derived from biodiesel amount to 440-580 €/tonne [12] and seem to have stabilised at this level. In
March 2007, the prices were around 400-450 €/tonne for tallow derived glycerin (99,5%) for delivery in
North West Europe (24-27 $ct/pound; [13]).
The crude glycerin market moves at levels around 0-150 €/tonne depending on a.o. the purity (50-
90%), water and residual methanol content. The refined glycerin market is described as being strong
(with new feed and chemical applications) while the crude glycerin market is described as weak [12].
The combination of high fossil oil prices and historically low glycerin prices have resulted in the
increased application of glycerin as an ideal platform chemical in the chemical and pharmaceutical
industry.
In Germany, it is allowed to both use crude and refined glycerin as a feed (pellet) ingredient [14]. The
quality demands for the crude glycerin are max. 0,5% methanol and min. 80% glycerin while it is
common to use only 2-5% of crude glycerin in the animal feed mix for poultry and pigs. One significant
drawback however, is that the biodiesel factory itself has to have GMP or GMP+ certification (most of
the biodiesel plants don’t have these certificates).
The existing biodiesel-derived crude glycerin is of poor quality and requires expensive refining before
it is suitable for new product technologies. Current glycerin refining technology requires significant
economies of scale to be economical.
The application of heterogeneous catalysts in biodiesel factories results in a much purer crude glycerin
and thus makes smaller scale and low-cost refining at the biodiesel plant viable.
Research on the enzymatic trans-esterification process for biodiesel production is still in an early
developmental stage, as this is still a relatively new field of study. Important studies in that field are
done by F. Yagiz, et. al [17] and Y. Shimada, et. al.[19]) while also a lot of progress was made by the
University of Cordoba [47]. Work at the University of Cordoba has shown that up to 100 re-uses of the
immobilised enzymes can take place and a PhD thesis will follow shortly. The well-known supplier of
ion exchange granulates Lanxess also investigated lipase immobilisation for biodiesel production and
claims a lifetime of 1 year using their Lewatit OC 1600 granulate as carrier material for the enzymes
[58 and below]. Lanxess gives several advantages for enzymatic biodiesel production and apparently
does not use any KOH.
The two studies by Y. Shimada, et. al. and F. Yagiz, et. al. have each focused on the use of a different
lipase for the trans-esterification process, namely Candida antarctica and Pseudornonas cepacia.
Many other lipases remained untested as yet, and will remain to be tested in the future as
development in the field matures.
There is general consensus in the studies that the enzymatic production of biodiesel is a superior
method as compared to conventional chemical trans-esterification, considering the lower complexity of
the reaction process and the absence of waste products, in particular soap (produced due to presence
of free fatty acids in the waste oil), which will create environmental problems if disposal is not handled
appropriately. As is shown in figure 3.1, also a cleaner biodiesel is produced. The enzymatically
produced crude biodiesel does contain less KOH and water washing of the biodiesel could thus be
substituted for a dry washing of the biodiesel. It has also been concluded that generally, the
immobilization of the lipases led to a higher activity of the lipases, which translates into a higher yield
of methyl esters (biodiesel) and a faster rate. However, the enzymatic trans-esterification process is
more time consuming than the chemical trans-esterification and the supercritical trans-esterification
process.
In enzymatic trans-esterification, although the reaction mechanism is similar to that of chemical trans-
esterification, an enzyme, specifically lipase, replaces the chemical catalyst. The lipase [19] catalyzes
hydrolysis1, esterification as well as trans-esterification.
Methanolysis (hydrolysis using methanol instead of water) of vegetable oil with a lipase is reported to
produce more effective results for the biodiesel production using waste oils. Enzymatic trans-
esterification offers a feasible option over the conventional method where the end product is not
1
Hydrolysis is the cleavage of an ester with water back to a carboxylic acid and alcohol
Types of lipase
Many laboratory experiments have been conducted to determine the effectiveness of enzymes in
replacement of chemical catalysts in trans-esterification production of biodiesel. The most common
and widely used enzyme in this initial trial stage is lipase, due to many of its favorable characteristics.
Lipases are water-soluble enzymes that catalyze the hydrolysis of ester bonds in water–insoluble lipid
substrates. They attack specific positions on the glycerol backbone of the bio-oils to allow the free fatty
acids to react with the alcohols more efficiently. Although there are many different lipases present
around, only a few selected lipases are selected for most of these experiments currently.
The Candida antarctica lipase [19] is a very common type of enzymes used in many experiments. It is
a moderately thermo-stable enzyme that is able to retain most of its activity for many hours when
incubated between 30 – 40°C. However, the enzymatic activity of the lipase is reduced when
temperature increases beyond 40 °C, exposed to higher water concentrations as well as xenobiotic2
compounds.
Another lipase used in the enzymatic trans-esterification of biodiesel is the Pseudornonas cepacia.
According to the experiments carried out by the University of Nebraska, Lincoln in 2004 [18], they
found that this strain of lipase resulted in the highest yield of alkyl esters produced. However, little
information has been known more extensively about the lipase as of now.
Although there are only two examples of lipases discussed here, there are still many other more
lipases suitable for the enzymatic trans-esterification of biodiesel production, but still currently
2
A xenobiotic substance is a foreign chemical not normally found in an organism and is in higher
than expected concentrations
In the third technique, the lipase is captured within a matrix of polymer. This method is preferred and
has already received considerable attention in recent years. This is because it better stabilizes the
lipase as compared to physical adsorption, and uses a simpler procedure than the covalent bonding
method. In general, the physical entrapment of lipase maintains the activity and stability of the
immobilized lipase.
The water, initially present in the waste oil, is transferred onto the polar glycerol and subsequently
removed from the system. Hence water concentration drops with repeated reactions, thereby
increasing methanolysis and hence the conversion rates. The diagram in fig. 3.2 shows the reaction in
cycles, coupled with a three-step methanol addition.
Alcohol concentration
As mentioned earlier, the insoluble methanol can inhibit the lipase activity, and is an irreversible
reaction, therefore resulting in a lower alkyl ester yield.
Figure 3-3 Methanolysis of vegetable oil with varying amounts of methanol using immobilized
Candida antarctica lipase. Conversion is expressed as the amount of methanol consumed.
This inactivation problem can be overcome by pre-treatment with higher order alcohols or the step-
wise addition of methanol into the mixture, as methanol is more soluble in alkyl esters than in oil [18].
In this way, there will be minimal methanol present to inhibit the enzyme activity. Results [19] have
shown that the Candida antarctica lipase can be used repeatedly for more than 100 days without
showing effects of inactivation using both three-step and two-step methanolysis.
Temperature
A joint research study [16] by 3 institutions in Turkey, Kocaeli University, University of Marmara and
the Marmara Research Center, showed the effects of temperature on lipase immobilized onto
On the other hand, for more practical and cost-effective reasons, immobilized lipases are usually used
in repeated batch processes due to their high costs. A continuation to the single step process
experiment carried out by the 3 institutions in Turkey, they investigated the activity of the enzymatic
activity at different temperatures.
According to their results, at 45°C the immobilized lipases displayed little or no loss of enzymatic
activity for the first 2 repeated cycles, and following 7 cycles, they retained about 36% of their initial
activity. At a higher temperature 55°C, the enzymatic activity was significantly different. Just after one
cycle, the activity of the immobilized lipases dropped almost 40%; and after 7 cycles, the remaining
activity was only 14% that of the initial activity.
From the above experiments, it is clear that for single processes, temperature does not have
significant effect on the activity of the lipases. However, after repeated usages, the higher the reaction
temperatures, the faster the rate of inactivity will occur. Thus, for practical reasons, temperatures
should be maintained at 45°C so that maximum activity of the lipases can be sustained.
pH
The 3 Turkish Institutions [16] further varied the pH values of the system to determine its effect on the
resulting yields. At different pH values, the amount of lipases chemically adsorbed onto the supporting
structure differs. Experiments showed that the highest lipase adsorption occurs at pH 8.5, and the
amount absorbed decreased more drastically when pH increases, than it decreases. As a rule of
thumb, enzymatic trans-esterification should not be operated at pH values higher than 8.5. However,
we should bear in mind that the results obtained from this literature are based on chemical adsorption
of lipase immobilization. Recently experiments at the University of Cordoba has shown that
immobilization methods through covalent bonding makes higher optimal pH values possible.
Although enzymatic trans-esterification produces a less contaminated biodiesel and by-product and is
considered a feasible alternative, the chemical catalyzed reaction is more commonly used and
recognized due to their shorter reaction time and lower overall costs [20].
As enzymes are costing at least 30 €/kg, it is necessary to re-use and immobilise the enzymes. The
University of Cordoba claims successful re-uses of the immobilised enzymes for up to 100 times while
a company like Lanxess claims a lifetime of up to 1 year while still retaining good enzyme activity.
More research is needed in order to develop cheap mass production of granulates containing
immobilised enzymes.
As previously mentioned, enzymes are very sensitive to the methanol concentration in the reaction
system; hence to ensure a minimum activity of the enzymes low concentrations of methanol should be
maintained at the initial stage of reaction. Some studies [20] have shown that it took 34 hours to
convert 97.3% of bio-oil in the refined vegetable oil to fatty acid methyl ester (FAME), in a two-step
methanol addition enzymatic process. When the refined vegetable oil is replaced by waste oil, the
conversion dropped to about 90.4%, taking a total of 48 hours.
The conventional trans-esterification is very sensitive to the purity of the bio-oil feedstock. Only well-
refined vegetable oil with less than 0,5-1,0% free fatty acid and less than 0,1% water can be used as
the feestock for conventional biodiesel production. The enzymatic process is also capable to convert
used cooking oil, waste animal fats and crude bio-oils [16, 47]. These lower quality feedstocks have
lower prices and thus the feedstock versatility may well be the most important economic advantage of
enzymatic biodiesel.
As was shown in table 2-1, conventional crude glycerin is polluted with 3,5-7 w-% of salt residue (ash).
Because the enzymatic process requires no homogeneous catalyst, only a high pH (8,5-11) the
amount of NaOH or KOH can be reduced by a factor of four compared to the conventional process.
Although significant progress has been made by a.o. the University of Cordoba, the trans-esterification
of bio-oils using lipases as catalysts is not completely without challenges. Two major challenges faced
by this alternative process are: the effects of the reactant [16] methanol and the product glycerol [18]
on the lipase activity. Both of the compounds lower the enzymatic performance and thus giving poorer
yields.
Solutions for these problems are given by the study conducted by D. Royon, et. Al [16]. He showed
that addition of some alcohols, which have 3 or more carbons, significantly increases the conversion
yields of biodiesels. When these alcohols are added in replacement of the methanol in the system, the
blank experiment showed no significant conversion of the oils to esters. Thus the higher alcohol
chains are not suitable substrates for the lipases used and do not interfere with the transesterification
process. Pretreatment of the enzymes with these higher alcohols reduces the inhibitory effects on the
enzymes of methanol and glycerol, because the higher alcohols increase the solubility of both reactant
methanol and product glycerol.
An example of a suitable higher order alcohol is the t-butanol (C4H9OH). Experimental results showed
that when sufficient t-butanol is present, an increase in methanol concentration increases the initial oil
consumption. The highest methyl-ester yield obtained occurs at a methanol-to-oil ratio of 3.6:1 in the
presence of t-butanol; when without this pre-treatment, any ratio greater than 1 inhibits lipase activity.
Some biodiesel equipment manufacturers like Axens and BDI are now experimenting with the
application of solid re-usable catalysts.
In the heterogeneous [15] transesterification process used by Axens the solid metal oxides such as
those of tin, magnesium, and zinc are known catalysts but they actually act according to a
homogeneous mechanism and end up as metal soaps or metal glycerates. In this new continuous
process, the transesterfication reaction is promoted by a completely heterogeneous catalyst. This
catalyst consists of a mixed oxide of zinc and aluminum, which promotes the transesterification
reaction without catalyst loss. The reaction is performed at higher temperature and pressure than
homogeneous catalysis processes, with an excess of methanol. This excess is removed by
vaporization and recycled to the process with fresh methanol.
Figure 3-4: Simplified [15] flow sheet of the new heterogeneous process
3.3 Conclusions
Preventing contamination of the glycerin fraction is the best way of reducing cost of purification of
glycerin and increasing the use of crude glycerin. Enzymatic biodiesel production with immobilized
enzymes looks very promising in this respect. Critical points of the enzymatic route are the relative
slow reaction compared to the conventional, homogeneous process and the number of times that the
enzymes can be used.
The heterogeneous catalytic process is also a promising way of producing glycerol fractions with high
purities. Critical points of this route are that the input oil has to be more pure and the stability of the
catalysts which has to sufficient justifying the higher investment costs.
Overall it can be concluded that it is possible to reach crude glycerin purities of 90-95% and thus
reducing purification costs.
There are different processes for refining glycerol. However, all of them involve soap splitting followed
by two main separation steps: salt removal and methanol removal. Some of the separation techniques
should involve vacuum because glycerol is a heat sensitive compound that splits into water and
decomposes at 180°C 3 .
Generally speaking, the following technologies may be used to further purify glycerin (after the soap
splitting step): fractional distillation, ion-exchange, adsorption, precipitation, extraction, crystallisation,
dialysis. The glycerin soap splitting followed by a combination of methanol recovery/drying, fractional
distillation, ion-exchange (zeolite or resins) and adsorption (active carbon powder) seems to be the
most common purification pathway.
Well-known companies who deliver crude glycerin purification plants are Desmetballestra and Buss-
SMS Canzler (ion exchange equipment). Chemical companies like Rohm & Haas and Lanxess supply
ion-exchange granulates while a company like Norit supplies powder and granulated activated carbon.
3
Brockmann, R., Jeromin, L., Johannisbauer, W., Meyer, H., Michel, O. and Plachenka, J.(1987).Glycerol distillation process.
US Patent No. 4,655,879
Figure 4-2 shows laboratory set-up for distilling the crude glycerin resulting in a 90-95% glycerin purity.
4
Knothe, G., van Gerpen, J. & Krahl, J. (2005). The biodiesel handbook. Illinois: AOCS Press.
5 R. Christoph, B.Schmidt, U.Steinberner, W. Dilla, R.Karinen. (2006). Glycerol, Ullmann’s Encyclopedia of Industrial Chemistry: electronic release, 6th ed.
John E. Aiken6 has made some improvements in glycerol purification process. He proposed five
separation steps, which can be conducted in either batch or continuous mode (Figure 4-5). This
process is claimed to be able to produce glycerol of higher than 99.5% purity from typical crude
glycerol, which contains a mixture of mono-, di- and triglycerides, excess methanol, water, fatty acid
alkyl esters, residual catalyst and salt.
6
Aiken, J.E. (2006). Purification of glycerin. US Patent No.7,126,032 B1
iii) Decanter
A decanter is placed after the second reactor. It serves as a feed tank for the flash distillation column
and a separator to remove oil layer of the glycerol stream by lowering the pH below 7 and skimming it
from the glycerol layer. The recycle stream from the bottom of the flash distillation column is mixed
with the glycerol stream in this tank.
Figure 4-5. Simplified flow sheet of the recent development process, based on US 7,126,032 B1
v) Adsorption columns
The last step of glycerol refining is the removal of colour and trace impurities. There are lots of
material that may be used as adsorbent, such as activated carbon, ion exchange resins and molecular
sieves. The purified glycerol is then pumped to a storage tank.
Activated carbon powder is, with surface areas between 500-1500 m2/g and sizes <150 micron, a very
suitable adsorption medium to adsorb organic molecules, but is rather expensive to regenerate the
carbon. Operational costs will be high when using a column adsorption, because of the high viscosity
of the crude glycerin and the high pressure drop. Activated carbon is applied because of its good
properties in waste water cleaning.
7
Megtec USA
The companies Rohm & Haas and Lanxess sell granular ion exchange resins which are also used for
glycerin purification (a.o. salts, colour and odour removal). But most important is the separation of
water and glycerin molecules based on affinity and particle size. Water molecules that are bound to
glycerin molecules are difficult to separate. It is therefore important to find a suitable type of adsorbent
with respect to high separation efficiency (resolution) at a high volume flow capacity and low pressure
drops. Next table summarises some chromatographic techniques with their resolutions and volume
flow capacities.
It should be possible to design new type of adsorbent media specific for glycerin that meet more or
less of the mentioned criteria. The first process set-up can be an ion exchange column with a second
column added. This column could be based on affinity properties adsorbent and (gel) permeation
principle. This principle is shown in the following figure:
8
Adapted from Biotol, Product Recovery in Bioprocess Technology (1992)
These characteristics are rather bad to apply on large scale, but new developments show potential for
higher fluxes at stable pressure drops, however these are far from economical feasible applications.
The figure below shows typical pressure drop calculations for continuous flow ion exchange granulate
columns [58].
Enzymatically produced crude glycerine will probably have a purity of 90-95%. Both soap splitting and
fractional distillation of the 90-95% glycerine can probably be eliminated. It seems likely that the use of
activated carbon powders and/or ion adsorption (zeolites or resins) techniques will probably be
sufficient to obtain a glycerine purity of 99,5%.
Regeneration of the granulates of activated carbon may typically be performed with methanol, acids,
steam or water.
The figure below shows a technical feasible column adsorption process on laboratory scale
(batch process). The resolution of the process is rather good but only possible at high retention times
and high pressure drop. Actual laboratory tests of column purification of enzymatic glycerine are
necessary to gain a further understanding about the process, possible absorbents (f.i. wood powder,
carbon, clay minerals), regeneration, retention times and purities that can be obtained.
This means that most of the costs are not involved in the energy but in the investment of equipment.
The dependent on the application and pricing of the glycerol it is interesting to invest in this system or
not. Cleaner crude glycerol can in this respect reduce the number of evaporating steps and thus
reduce the purification steps.
The glycerine applications and market prices in dependence of purity are described in chapter 2.
A challenge faced for enzymatic trans-esterification is the high costs associated with lipases as
catalysts [16]. Much progress has been made with the successful immobilization of the enzymes and
re-use up to 100 times [47]. The well-known supplier of ion exchange granulates Lanxess also
investigated lipase immobilisation for biodiesel production and claims a lifetime of 1 year using their
Lewatit OC 1600 granulate as carrier material for the enzymes [58].
Enzyme costs depend strongly on their purity starting from 30 €/kg up to more than 10.000 €/kg. For
enzymatic biodiesel production it is not necessary to use high grade expensive enzymes [47] and thus
a price of 100 €/kg of enzymes (including immobilisation) was taken. In contrast, KOH costs only 0.6
€/kg.
For this economic evaluation it was assumed the enzymatic process reduces the amount of KOH used
with a factor of 4 (from f.i. 1,8 wt%/kg oil to 0,45 wt%/kg oil). This may be conservative. The
consumption of enzymes was set at 0,1 wt % of immobilised enzymes/ kg oil and 100 re-uses. In table
5.1 it was conservatively assumed that glycerine purification costs for enzymatically produced glycerin
(input 90-95% purity) are 100 €/tonne while for conventionally produced glycerine (75%-85% purity)
this was assumed to be 150 €/tonne. The difference in purification costs follows from the elimination of
a soap splitting and fractional distillation step when purifying enzymatically produced glycerine.
Table 5.1 shows that enzymatic biodiesel production may easily result in significantly lower operating
costs (excl. pure plant oil purchases). Table 5.1 only aims to display the main differences (differential
costs) between the conventional and enzymatic biodiesel process.
From table 5.1 it may also be concluded that the biggest economic advantage when applying
enzymatic biodiesel production processes may not result from decreased glycerine purification costs
but from drastically decreased feedstock oil costs. As the enzymes are able to also convert FFA’s and
Cash flows for bio-oil purchasing Cash flows for bio-oil purchasing
Bio-oil purchasing 71.428.571 €/yr Bio-oil purchasing 61.224.490 €/yr
The following processes can be utilized in obtaining useful derivatives from glycerol:
Esterification, Etherification, Oxidation, Reduction, Amination, Halogenation, Phosphorylation,
Nitration and Sulfaction.
The complete schematic flow sheet is given for the production [36] [40] and utilization routes for glycerin
(R&D Potential for biodiesel, NREL 2003).
• Reduction
The specialty chemicals from glycerin are due the fact that Glycerol provides a C3 building block for
complex structures. It is easily modified by reacting –OH functional groups and it can produce water
soluble, nontoxic, and nonflammable products.
The glycerin [2] recovered from the transesterification reaction is etherified with methanol, ethanol or
butanol using another proprietary heterogeneous catalyst. The former methanol plant at Delfzijl
acquired in 2006 from the joint owners DSM, Akzo Nobel and Dynea by BioMethanol Chemie Holding
(a consortium of Ecoconcern, the NOM, the investor OakInvest, and the process technologists Sieb
Doorn and Paul Hamm) produces fossil methanol and plans to use glycerin as a raw material for
producing bio-methanol. The bio-methanol is intended for use in the first instance as a petrol additive
but at a later stage it could power fuel cells. The plant formerly produced methanol from natural gas
but was closed down because this process was no longer profitable. The plant [1] will use a new
process to make bio-methanol from glycerin.
For the biodiesel process this can eliminate the role of methanol:
Virent [26] has developed the novel APR (Aqueous-Phase Reforming) process and has shown that it is
effective for generating hydrogen from aqueous solutions of glycerol. The APR process is a simple
one-step reforming process that can generate easily purified hydrogen and as such is especially cost
effective.
Raw glycerol is refined to remove contaminants such as KOH and alcohols, and the resultant pure
material is used in many applications including food and personal products. Raw glycerol can be
mixed with water and the resulting aqueous solution can be fed to the APR process that generates
hydrogen in a single reactor. The effluent gas from the APR process can be purified to produce high
purity hydrogen.
The APR process generates hydrogen by reacting a carbohydrate, in this case glycerol, with water to
form carbon dioxide and hydrogen as follows:
APR process runs at low temperature, an alternative method of providing process energy will be to
utilize waste heat streams from other associated processes. Thermal efficiencies of the process can
be maintained via proper heat exchange (i.e. preheating feed to the reactor by exchanging with the
reactor effluent). Alkanes such as methane, ethane, and propane are also formed in low
concentrations in the APR reactor. The alkane formation is an exothermic process, and while alkane
formation lowers the hydrogen yield, the formation of these compounds provides heat for the
endothermic hydrogen generation process.
The non-condensable gas stream leaving the APR contains predominately CO2 and H2. Hydrogen can
easily be purified from this gas stream utilizing pressure swing adsorption (PSA) technology.
Importantly, the gas stream that exits the APR is at desired feed pressures for the PSA unit (between
16 and 40 bar). Accordingly, the PSA unit does not need an expensive and energy consuming
compressor to provide the necessary feed pressure. This results lower capital costs and increased
system energy efficiency. Another important feature is that the PSA technology will generate a waste
hydrogen stream (typically 10 to 20 percent of the feed) due to the pressure swing and purging cycles.
This waste stream would also contain the alkanes produced in the APR process. Combustion of the
waste hydrogen and alkanes would provide much of the necessary processing heat for the reactor.
Finally, the waste stream could provide a starting reactant for the production of biodiesel. Hydrogen
and CO2 reacted over a catalyst of copper and zinc is converted to methanol by the reaction
A comparable steam reformer utilizing non-renewable natural gas is expected to be 56% efficient and
could generate hydrogen at a cost of 2.66 euros per kg of hydrogen (National Research Council,
2004). Furthermore, a comparable unit that generates hydrogen via the electrolysis of water would
generate hydrogen at a cost of 4.94 euros per kg of hydrogen (1 kg equals a heating value of 121 MJ).
.
The microbial [25] conversion of glycerol to various compounds has been investigated recently with
focus on the production of H2 and ethanol from glycerol. H2 is expected to be a future clean energy
Glycerol [30] is first adsorbed and dehydrogenated reversibly on the metal catalyst to form
glyceraldehyde. The glyceraldehyde then desorbs from the catalyst and can react through four
different paths in the basic media: the retro-aldol mechanism to form the precursor of ethylene glycol
(glycolaldehyde), oxidation and subsequent decarboxylation to also form glycol aldehyde, dehydration
to the precursor of propylene glycol (2-hydroxypropionaldehyde) or degradation to unwanted side
products. The two glycol precursors could potentially also degrade to unwanted side products. Finally,
the respective glycol precursors are hydrogenated by the metal function to the product glycols.
The flow scheme below gives the reaction pathway for the production of glycols from glycerol.
Experimental results of the decomposition of glycerol in near and supercritical water are presented
considering measurements in the temperature range of 622–748 K, at pressures of 25, 35, or 45 MPa,
reaction times from 32 to 165 s, and different initial concentrations. The reaction was carried out in a
tubular reactor and a conversion between 0.4 and 31% was observed. The main products of the
glycerol degradation [42] are methanol, acetaldehyde, propionaldehyde, acrolein, allyl alcohol, ethanol,
formaldehyde, carbon monoxide, carbon dioxide, and hydrogen.
The fact that the measured composition of the product mixture at constant temperature is depended
on the density was taken as an indication, that these products could be formed by competing ionic and
free radical reaction pathways. Usually in gas kinetics, the product composition changes with
temperature. This is due to the different activation energies, the concentration effect on bimolecular
elementary reaction steps and in a minor extent with pressure. In water, the drastic dependence on
pressure is likely a consequence of the competition between reactions with different polarity.
Formaldehyde: In the ionic mechanism is formed by the same reactions as acetaldehyde. In the
radical mechanism, nearly all reaction paths, in which a decomposition of C3-radicals to C1 and C2-
fragments take place, lead to CH2O formation. Formaldehyde is here only an intermediate product,
because it is oxidized to CO or CO2 by a sequence of reactions with OH-radicals.
Carbon monoxide: Is formed by the reaction of CH2O with an OH-radical to water and the CHO-
radical, which consecutively decomposes to CO and H-atom.
Carbon dioxide: Is formed by oxidation of CO with OH to form CO2 and an H-atom.
Hydrogen: Is formed by all metathesis reactions (mostly with glycerol) of the H-atoms. The H-atoms
can also react with the OH-groups of a substance (mostly glycerol) to form water and a radical.
Propionaldehyde: Is a product measured only at low concentration. There is no formation in a simple
ionic pathway imaginable. The radical pathway to propionaldehyde is also rather complicated. One of
the paths starts from allyl alcohol (which can be considered as an isomer of propionaldehyde). A
radical addition (e.g. H-atom) followed by a combination of radical isomerization (enol-type) and
radical elimination can yield the propionaldehyde. A lot of other minor products were measured during
the experiments, which are only partially included in the reaction mechanism: ethanol, acetone,
ethane, ethene, propene, propane, butenes, butanes, methyl-hydroxy-dioxanes and other products of
higher molar masses. Most of the minor products are only found at higher temperatures and are most
likely formed via radical reaction pathways decomposing the main products.
The method presented here may allow for economic operation of a small-scale Fischer–Tropsch [45]
reactor by producing an undiluted H2/CO gas mixture. The method reduces the capital cost of the
Fischer–Tropsch plant by eliminating the O2 plant or biomass gasifier and subsequent gas-cleaning
steps. The conversion of glycerol into CO and H2 takes place by Equation (1).
The endothermic enthalpy change of this reaction (350 kJ/mol) corresponds to about 24% of the
heating value of the glycerol (1480 kJ/mol). The heat generated by Fischer–Tropsch conversion of the
CO and H2 to liquid alkanes such as octane (412 kJ/mol) corresponds to about 28% of the heating
value of the glycerol. Thus, combining these two reactions results in the following exothermic process,
with an enthalpy change (63 kJ/mol) that is about 4% of the heating value of the glycerol:
Catalysts consisting of Pt supported on Al2O3, ZrO2, CeO2/ZrO2, and MgO/ZrO2 exhibited deactivation
during time-on-stream, whereas the Pt/C catalyst showed stable conversion of glycerol into synthesis
gas for at least 30 hours. The catalyst with the most acidic support, Pt/Al2O3, showed a period of
apparently stable catalytic activity, followed by a period of rapid catalyst deactivation. The reactor
initially operates at 100% conversion, glycerol is present only in the upstream portion of the catalyst
bed in the tubular reactor and a deactivation front moves from the reactor inlet to the outlet as olefinic
species are formed from glycerol on acid sites associated with alumina, followed by deposition of coke
from these species on the Pt surface sites. The most basic catalyst support, MgO/ZrO2, showed rapid
deactivation for all times-on stream. The most stable oxide-supported catalyst appears to be Pt on
CeO2/ZrO2; however, the performance of this catalyst is inferior to that of Pt supported on carbon. The
different deactivation profiles displayed in Figure 8 for the various catalysts suggest that the support
plays an important role in the deactivation process. Figure 8d shows the rate of formation of C2-
hydrocarbons (ethane and ethylene) normalized to the rate of H2 production for the various supported
Pt catalysts. Negligible amounts of C2-hydrocarbons were formed on the Pt/C catalyst. In contrast,
catalysts consisting of Pt supported on the various oxides formed measurable amounts of C2-
hydrocarbons, and the C2-TOF/H2-TOF ratio (TOF=turnover frequency) increased with time-on-
Figure 6-5 Performance [45] of supported Pt catalysts with Variation with time-on-stream.
For these studies of reaction kinetics, 0.060 g of 5 wt% Pt/C was used. [a] Glycerol feed 30 wt%, 623
K, 1 bar. [b] Feed flow rate 0.32 cm3/min, 623 K, 1 bar. [c] Point taken after 2 h time-on-stream. [d]
Glycerol feed 30 wt%, 0.32 cm3/min and 1 bar.
The catalytic conversion of polyols to H2, CO2, and CO involves the preferential cleavage of C-C
bonds as opposed to C-O bonds and Pt-based catalysts are particularly active and selective for this
process. Under these reaction conditions, the surface is covered primarily by adsorbed CO species. A
strategy for a catalyst that converts polyols into synthesis gas and is active at low temperatures is to
facilitate the desorption of CO, thereby suppressing the subsequent WGS step and improving the
turnover of the catalytic cycle by regenerating vacant surface sites. Accordingly, we require materials
that possess the catalytic properties of Pt with respect to selective cleavage of C-C versus C-O bonds,
but that have less exothermic enthalpy changes for CO adsorption; Pt–Ru and Pt–Re alloy catalysts fit
this description. These results demonstrate that the conversion of glycerol to synthesis gas can be
accomplished at temperatures well within the ranges employed for Fischer–Tropsch and methanol
syntheses, thus allowing for the efficient combination of these processes at low-temperature catalytic
route for converting glycerol into H2/CO gas mixtures that are suitable for combination with Fischer–
Tropsch and methanol syntheses.
Glycerol [32] can be esterified to polyglycerols and especially polyglycerols-esters (PGEs) are gaining
prominence. Esterification of glycerol could be selective to monoglycerides over cationic resins. Never
the less, polyglycerols and polyglycerols esters as well as acrolein were obtained as main by-
products. The schematic representation of the etherification of glycerol to poly glycerols is given
below.
Glycerol etherification is carried out at 533K in a batch reactor at atmospheric pressure under N2 in the
presence of 2 wt% of catalyst; water being eliminated and collected using a Dean-Stark system.
Reagents and products are analysed with a GPC after silylation (It involves the replacement of an
acidic hydrogen on the compound with an alkylsilyl group).
Conventional [33] glycerolysis requires high temperatures (220–260ºC) to increase the solubility of
glycerol in the fat phase, the addition of nitrogen gas to prevent oxidation and the presence of an
inorganic catalyst. The reactants must also be vigorously stirred throughout the reaction and, at the
end of the reaction, the catalyst must be neutralized and reaction mixture must be rapidly cooled to
prevent reversion. Conducting glycerolysis in supercritical carbon dioxide (SC-CO2) simplifies the
conventional process. Under ambient conditions, oil and glycerol are immiscible and the main reason
for conducting glycerolysis reactions at 250ºC is to increase the solubility of glycerol in oil. With the
Hydrogenolysis [28] of glycerol to propylene glycol was performed using nickel, palladium, platinum,
copper, and copper-chromite catalysts. The effects of temperature, hydrogen pressure, initial water
content, choice of catalyst, catalyst reduction temperature and the amount of catalyst were evaluated
by the authors. At temperatures above 200 ºC and hydrogen pressure of 200 psi, the selectivity to
propylene glycol decreased due to excessive hydrogenolysis of the propylene glycol. At 200 psi and
200 ºC the pressures and temperaures were significantly lower than those reported in the literature
while maintaining high selectivities and good conversions. The yield of propylene glycol increased with
decreasing water content. Propylene glycol, i.e. 1,2 propanediol, is a three-carbon diol with a
steriogenic center at the central carbon atom. Propylene glycol is a major commodity chemical with an
annual production of over 1 billion pounds in the United States and sells for about 0.53 euro cent per
pound with a 4% growth in the market size annually. The commercial route to produce propylene
glycol is by the hydration of propylene oxide derived from propylene by either the chlorohydrin process
or the hydroperoxide process. There are several routes to propylene glycol from renewable
feedstocks. The most common route of production is through hydrogenolysis of sugars or sugar
alcohols at high temperatures and pressures in the presence of a metal catalyst producing propylene
glycol and other lower polyols. The summary of the overall reaction of converting glycerol to propylene
and ethylene glycols is given below:
Various noble metals (Ru/C, Rh/C, Pt/C, and Pd/C) and acid catalysts [an ion-exchange resin
(Amberlyst), H2SO4 (aq), and HCl(aq)], the combination[29] of Ru/C with Amberlyst is effective in the
dehydration and hydrogenation (i.e. hydrogenolysis) of glycerol under mild reaction conditions (393 K,
8.0 MPa). The dehydration of glycerol to acetol is catalyzed by the acid catalysts. The subsequent
hydrogenation of acetol on the metal catalysts gives 1,2-propanediol. The activity of the metal catalyst
and Amberlyst in glycerol hydrogenolysis can be related to that of acetol hydrogenation over the metal
catalysts. Regarding acid catalysts, H2SO4 (aq) shows lower glycerol dehydration activity than
Amberlyst, and HCl(aq) strongly decreases the activity of acetol hydrogenation on Ru/C. In addition,
the OH group on Ru/C can also catalyze the dehydration of glycerol to 3-hydroxypropionaldehyde,
which can then be converted to 1,3-propanediol through subsequent hydrogenation and other
degradation products.
The role of OH species on Ru is thought to be important because Ru/C is much more active than other
noble metal catalysts in glycerol hydrogenolysis Another important point is that Ru species can
catalyze dehydration to 3-hydroxypropionaldehyde, although two dehydration routes can be traced to
3-hydroxypropionaldehyde and acetol. When this OH species attacks H linked to terminal carbons, 3-
hydroxypropionaldehyde is produced, which can explain the dehydration selectivity. However, the
reason why OH species do not attack H linked to center carbons remains unclear:
In the case of the Amberlyst, the active species is a proton. Acetol is formed when the proton attacks
OH linked to terminal carbons:
Given below is the reaction scheme of glycerol hydrogenolysis and degradation reactions.
[36]
In the flow scheme below, the flow scheme describes the conversion of glycerin and isobutlylene
on acid catalysis to give mono, di and tri glycerols.
In figure 9, the selective analysis of GTBE: diesel ratio is given with isobutene and glycerol as
reactants. The processing costs, NOx reduction and diether selectivity is shown. The variation
percentage of the diesel cost decreases as the use GTBE increases, the NOx however remains more
or less the same.
Figure 6-6 Selective analysis [36] GTBE/diesel ratio for isobutene and glycerol as reactants
In the glycerol ethers synthesis [38], the ethers are excellent oxygen additives for diesel fuel.
Oxygenated diesel fuels are of importance to both environmental compliance and efficiency of diesel
engines. A number of studies on the preparation of glycerol ethers by using different catalytic systems
have been reported. The reaction can be carried out with homogenous or heterogeneous acidic
catalysts. Recently, we developed a procedure of catalytic synthesis of high value glycerol ethers
(primarily di and tri-tert-butyl), obtained directly from glycerol and isobutene contained in the cracking
derived fraction. Several products were obtained in this reaction, the desired ones being: 1,3-di-tert-
butoxy-propan-2-ol (2a), 2,3-di-tert-butoxypropan- 1-ol (2b), 1,2,3-tri-tert-butoxy-propane (3). Efficacy
of 2a, 2b, and 3 for biodiesel fuel results from their decrease of emission of particulate matter,
viscosity, cold filter plugging point, and cloud point. Given below is the flow scheme for tert-Butylation
of glycerol catalysed by ion-exchange resins.
The best [39] results of glycerol tert-butylation by isobutylene at 100% conversion of glycerol with
selectivity to di- and tri-ethers larger than 92% were obtained over strong acid macro reticular ion-
exchange resins. Di- and tri-tert-butyl ethers of glycerol are potential oxygenates to diesel fuel. There
are known some possibilities improving burning characteristics of diesel fuels with oxygenate
The etherification of glycerol is preferred on primary hydroxyl groups (formation of 1-tert-butyl glycerol
and 1,3- di-tert-butyl glycerol). Di- and tri-tert-butyl ethers of glycerol are usable as potential oxygenate
additives to diesel fuels because of their blending with diesel. Mono-tertbutyl ether of glycerol (MTBG)
has a low solubility in diesel fuel and therefore the etherification of glycerol must be directed to the
maximum formation of di- and tri-ethers. The etherification of glycerol with isobutylene or tertbutyl
alcohol using strong acid ion-exchange resins amberlyst type and two large-pore zeolites H-Y and H
Beta was used. The highest glycerol conversion of 100% was obtained over strong acid macro
reticular ion-exchange resin A 35 at 60ºC. Higher temperature (90 ºC) causes considerable drop in
conversion and yield of desired di- and tri-ethers mainly in the case of acid ion-exchange resins.
6.12 Conclusions
Glycerin is one of the oldest chemicals and the possibilities of use are numerous. For most of the
applications glycerin has to be pure enough in order not to contaminate a catalyst or bacteria. Which
applications are most promising depends on technical and economical criteria in combination with
environmental benefits.
Promising applications from a market point of view are methanol, hydrogen, ethanol, 1,3 propanediol,
propylene glycol and GTBE. From a technical point of view these chemicals can all be made. The
One process looks especially useful for glycerin arising from conventional biodiesel production:
making hydrogen via the aqueous phase reforming (APR) process. The APR system generates
hydrogen from aqueous solutions of oxygenated compounds, such as biomass-derived glycerin, in a
single-step reactor process. Sodium hydroxide, methanol and the high pH levels common in low-grade
crude glycerin actually help the process. The producer claims that approximately 5 kg of glycerin can
be converted to 0.75 kg of hydrogen (50% efficiency). With the cheap crude glycerol there is a
possibility to generate gas from glycerol for less than 2 euro per kilogram.
Table 5-3 Indicator score for feasibility of different chemicals from glycerol.
Development Price
9
Market
phase Euro/ton
Methanol D/C ο 250
Hydrogen R ++ 2200
Ethanol R ++ 740
1,3 propane diol D ++ 1000
propylene glycol C ++ 1500
GTBE D ο 750
Polyglycerols D ο 1000
Conclusion is that all above mentioned chemicals have a great market potential based on glycerol
chemistry.
9
C= Commercial, R=Research, D = Demonstration
Significant progress has been made by researchers at especially University of Cordoba towards the
development of the enzymatic biodiesel process. One doctoral thesis will be published shortly and an
enzymatic pilot plant will be erected in Spain. Enzymatic biodiesel production has the following
advantages:
-not sensitive to lower oil qualities (FFA and water content);
-much purer glycerin (90-95%) not spoiled by catalyst;
-operation at lower temperatures / better energy balance;
-lower chemical catalyst cost;
-much lower glycerin purification costs
-lower biodiesel purification costs.
Drawbacks which can be mentioned are the sensitivity of the process towards reaction conditions
(optimal process step sequences, degradation of enzymes, methanol-enzyme interaction) and the
economic necessity for (cheap) immobilisation and many re-uses of the enzymes. Although some
researchers claim 100 times re-use and/or lifetimes of 1 year much more independent lab-work is
needed to prove the viability of the enzymatic process.
Glycerin purification
The application of heterogeneous catalysts in biodiesel factories results in a much purer crude glycerin
and thus makes smaller scale and low-cost refining at the biodiesel plant viable. It is expected that
enzymatically produced crude glycerine could result in a purity of 90-95% when compared to the
conventional 75-85% purity. This is the result of the much lower KOH quantities used and the much
lower soap concentrations due to the fact that free fatty acids are also converted towards biodiesel.
Both soap splitting and fractional distillation of the 90-95% glycerine can probably be eliminated. It
seems likely that the use of activated carbon powders and/or ion adsorption (zeolites or resins)
techniques will probably be sufficient to obtain a glycerine purity of 99,5%. However, it is highly
recommended to perform actual laboratory tests of column purification of enzymatic glycerine are
necessary to gain a further understanding about the process, possible absorbents (f.i. wood powder,
Economic evaluation
The application of enzymatic biodiesel will lead to lower chemical catalyst and much lower glycerine
purification costs. However, our evaluation has shown that the biggest economic advantage when
applying enzymatic biodiesel production processes may not result from decreased glycerine
purification costs but from drastically decreased feedstock oil costs. Attractive conversion routes
towards high value chemicals from glycerin have been identified. Many of these conversion routes
seem economically and technically feasible.
Glycerin is one of the oldest chemicals and the possibilities of use are numerous. For most of the
applications glycerin has to be pure enough in order not to contaminate a catalyst or bacteria. One
process looks especially useful for glycerin arising from conventional biodiesel production: making
hydrogen via the aqueous phase reforming (APR) process. The APR system generates hydrogen from
aqueous solutions of oxygenated compounds, such as biomass-derived glycerin, in a single-step
reactor process. Sodium hydroxide, methanol and the high pH levels common in low-grade crude
glycerin actually help the process. The producer claims that approximately 5 kg of glycerin can be
converted to 0.75 kg of hydrogen (50% efficiency). With the cheap crude glycerol there is a possibility
to generate gas from glycerol for less than 2 € /kg.
It is clear from the research results that the purification of glycerol produced through normal process
routes needs several steps and is energy consuming. Therefore the current biodiesel research is
focused on developing new ways of producing biodiesel without using base/acid catalyst which can
greatly reduce the downstream processing step.
This report shows that enzymatically produced biodiesel offers huge opportunities which result from
the much lower oil feedstock costs, decreased energy consumption and decreased glycerine
purification costs. We encourage that more research activities are directed towards the development
and commercialisation of the enzymatic biodiesel route.
Another new biodiesel process uses a two-step supercritical reaction process with adsorption refining
[55]. In this process reaction is carried out at a temperature greater than the critical temperature of
methanol without using base/acid catalyst. Excess methanol is used and fats with any amount of free
fatty acid content can be a raw material for the process. The reaction is carried out in two steps since it
is the most economical way of meeting the energy and pumping requirement of the process [55]. In
contrast to the conventional way of purifying the glycerol stream, the glycerol stream from the reactors
is treated in adsorption beds [55] which later on can be recovered by flashing it with methanol stream
and recycling it back to the reactor. Another process that is tried involves immobilized enzyme
catalysis. In this regard lipase catalyst has been used and it requires the lowest temperature condition
for the reaction and requires less equipment in the purification stage as compared to acid and base
catalysts [48]. The main bottleneck for applying enzymatic production is the cost of catalysts which
makes the process economically less appealing [2]. Another important commercially developed
(Esterfif-HTM process) biodiesel production [45] which is based on heterogeneous catalysis by mixed
oxide of zinc and aluminium allowed reaching glycerol purity level of more than 98% from
transesterification reaction. The reaction in this case is carried out in two successive stages with
excess methanol recycled to the reactor by evaporation. Glycerol is also removed continuously which
favours the forward transesterification reaction. The other important process that has been developed
is microwave irradiation production of biodiesel [49, 56]. The application of microwave energy
selectively energize polar molecules over non-polar and neutral ones thereby enhances selectively the
physical and chemical processes to biodiesel production. The conversion is almost 100% resulting in