You are on page 1of 21

ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

1
Triazine Chemistry: Removing H
2
S and Mercaptans
1


Thomas R. Owens and Peter D. Clark

Alberta Sulphur Research Ltd., University Research Centre, #6-3535 Research Road
N.W., Calgary, Alberta, Canada, T2L 2K8 [e-mail: pdclark@ucalgary.ca]


Abstract

Arguably, one of the most common methods for removing low levels of hydrogen
sulfide from natural gas (i.e., < 1 ton of sulfur per day) is triazine based chemical
scavengers. Triazines are produced from the reaction of primary amines such as
monoethanolamine or methylamine with aldehydes, specifically, formaldehyde. Despite
their wide spread usage throughout the industry only a few studies have been published
regarding the reaction of triazines with hydrogen sulfide. In studies conducted by Bakke
and coworkers, the kinetics and reaction products (i.e., dithiazines) for the reaction of
different triazines with hydrogen sulfide were investigated under various pH
conditions.
1,2


Another type of sulfur compound that intermittently needs to be removed from
hydrocarbon streams is mercaptans. Although some in the industry have indicated that
triazines can be used to remove mercaptans, there is no specific information in the open
literature regarding the reaction of triazines with mercaptans in regards to reaction
products or kinetics.

In this paper, we present a thorough background into the chemistry of triazines, in
particular, their reaction with hydrogen sulfide and potential problems produced from the
formation of dithiazine. In addition, the current results of a systematic study for the
reaction of two triazines, 1,3,5-tris(2-hydroxyethyl)hexahydro-s-triazine (1) and 1,3,5-
trimethylhexahydro-s-triazine (2) with different mercaptans: methanethiol (MeSH),
ethanethiol (EtSH) and n-propanethiol (n-PrSH) will be presented.



1
This paper will be presented at the Laurence Reid Gas Conditioning Conference, February 20-23, 2011.
ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

2
N N
N
OH
OH
HO
N N
N
1 2


Introduction

Hydrogen sulfide, a toxic and odorous gas, is present in many different natural
gas fields and many other hydrocarbon streams. The removal of large quantities of
hydrogen sulfide (i.e., 10-4000 ton/day of sulfur) is accomplished using the modified
Claus process. Mid level sulfur recovery, in the range of 1-10 ton/day, has been
problematic but has been done using a number of different processes such as LO-CAT

,
SulFerox

, CrystalSulf

and Shell Paques/THIOPAQ


TM
, just to name a few. Sulfur
capture below 1 ton/day is accomplished using chemical scavengers such as iron-based
absorbents, zinc oxides, porous carbon and solutions of hexahydrotriazines or simply
triazines.

Arguably, triazines are the most common chemical scavenger used in the field for the
removal of low level hydrogen sulfide from hydrocarbon streams. Triazines are compounds
which contain three carbon and three nitrogen atoms in a six-membered ring and can be
either carbon- or nitrogen-substituted. The functional group can have major affects on
the reactivity of a triazine; in fact, triazines in addition to being useful as chemical
scavengers also are used widely as herbicides, biocides and explosives. Despite the large
possible variation in triazine functionality, the most common triazine for removing
hydrogen sulfide is 1,3,5-tris(2-hydroxyethyl)hexahydro-s-triazine (1) primarily due to
the low cost of the raw materials, monoethanolamine and formaldehyde. The second
most commonly utilized triazine is 1,3,5-trimethyl-hexahydro-s-triazine (2). Our
discussions will be confined to 1 and 2.

The reaction between three moles of formaldehyde with three moles of either
monoethanolamine or methylamine produces three moles of water and one mole of 1 or
2, respectively. The synthesis is illustrated below in equation 1. The reaction is
exothermic and although no kinetic or thermodynamic data for the formation of the
triazine have been measured to the best of our knowledge, it is suspected that the reaction
occurs at or close to the limits of diffusion.


ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

3
N N
N
R
R R
RNH
2
H H
O
1 R = HOCH
2
CH
2
2 R = Me
+
(1) 3 3
+
3 H
2
O


Prior to discussing the reaction of 1 and 2 with hydrogen sulfide, there are two
important side reactions, hydrolysis and thermal degradation, that will be discussed since
both types of reactions can adversely affect the performance of triazine scavengers. It
should be noted, carbon dioxide does not react with triazines thus the presence of carbon
dioxide in the gas stream does not inhibit the scavenging efficiency of triazine solutions.
Another potential reactant is oxygen but to the best of our knowledge nothing has been
published in the open literature regarding this possible reaction. It is suspected that
triazines will undergo decomposition when left in contact with air for prolonged periods,
leading to possible deactivation and loss of scavenging ability.

Hydrolysis is the acid catalyzed reaction of 1 or 2 with water (equation 2) to
produce one molar equivalent of 1,3,5-trioxane and three molar equivalents of
monoethanolamine or methylamine, respectively. Unfortunately, 1,3,5-trioxane does not
react with either the free amines or with hydrogen sulfide at temperatures below 90C;
thus hydrolysis is an unproductive chemical reaction which only serves to diminish the
effectiveness of triazine solutions. Bakke and coworkers measured the rate of hydrolysis
of 1 and 2 in various pH buffered solutions.
1
These authors determined that hydrolysis
only became significant at pH < 10 and that 2 was approximately 20 times more stable to
hydrolysis than 1. It is important to note, that the work by Bakke and co-workers was
conducted with dilute solutions (0.10% triazine by volume) and the rate of hydrolysis in
concentrated triazine solutions, such as those used for hydrogen sulfide scavenging, will
be slower due to a lower activity of water. None the less, their work indicates that
aqueous triazine solutions are incompatible with acidic solutions.


N N
N
R
R R
O O
O
1 R = HOCH
2
CH
2
2 R = Me
(2)
H
2
O
pH < 10
+
3 RNH
2



In addition to hydrolysis, 1 can thermally decompose when it is exposed to high
temperatures. The reaction is an intramolecular cyclization of the 2-hydroxyethyl
ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

4
substituent to produce oxazolidine (equation 3); in this reaction one mole of triazine will
produce three moles of oxazolidine. This reaction has proved challenging for
quantitatively measuring 1 with chromatographic methods such as LC-MS
3
and GC-MC
4

which both expose the triazine to temperatures in excess of T = 200C for brief periods of
time. The utilization of chromatographic methodology requires either complexing 1 with
sodium ions for LC-MS
3
or derivatizing the hydroxyl groups with trifluoroacetyl groups
for GC-MS
4
. It is not possible to comment on long term degradation of triazine solutions
at lower temperatures since the rates of thermal degradation at various temperatures have
not been published. However, it is advisable to minimize either the storage time or the
storage temperature for fresh chemical in order to maintain the highest H
2
S capacity
possible. The thermal degradation reaction also suggests that formulations blended at
higher temperatures may have lower hydrogen sulfide capacities relative to those blended
at lower temperatures; it should be reiterated, that the reaction of MEA and formaldehyde
is an exothermic reaction.

N N
N
OH
OH
HO
N O
(3)
Heat
3
oxazolidine

The reaction of triazines with hydrogen sulfide (Scheme 1) occurs stepwise by
first producing thiadiazines 3 and 4 followed by the reaction with a second molecule of
hydrogen sulfide to produce 5 and 6, commonly referred to as dithiazines.
1,2
The
stoichiometry of the reaction is two molar equivalents of hydrogen sulfide reacts with one
molar equivalent of either 1 or 2 to produce two molar equivalents of amine
(monoethanolamine or methylamine) and one molar equivalent of 5-(2-hydroxyethyl-)-
1,3,5-dithiazine (5) or 5-methyl-1,3,5-dithiazine (6), respectively. In field applications,
only dithiazines are observed as reaction products which suggest that under field
conditions triazines and thiadiazines have similar reaction rates with hydrogen sulfide. It
is important to note, that dithiazines do not react with hydrogen sulfide under conditions
observed in the field; this reaction, if it were to occur, would produce 1,3,5-trithiane.










ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

5
Scheme 1


N N
N
R
R R
N N
S
R R
RNH
2

N S
S
R
RNH
2
H
2
S
+
H
2
S
+
3 R = HOCH
2
CH
2
4 R = Me
1 R = HOCH
2
CH
2
2 R = Me
5 R = HOCH
2
CH
2
6 R = Me


The rate constants for the reaction of triazines with hydrogen sulfide were
measured as a function of pH by Bakke and coworkers who determined that decreasing
the pH from 11.0 to 10.0 increased the reaction rate by a factor of ten for 1 and 2.
2
It is
interesting to note, that 1 was found to react thirty times faster with hydrosulfide than 2 in
laboratory experiments conducted with dilute triazine solutions and sodium hydrosulfide.
No explanation for the difference in reactivity between the two triazines was offered.
Unfortunately, the rate data cannot be applied directly to field applications since the data
were obtained under conditions which exclude mass transfer limitations (i.e., hydrogen
sulfide in the field must dissolve into the scavenger). However, the investigation does
indicate that the reaction is acid catalyzed and that initial steps in the reaction occur via
protonation of the triazine. The reaction mechanism most consistent with the available
data is illustrated in Scheme 2.

ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

6
Scheme 2

N N
N
R
R R
N N
N
R
R R
H
N N
R R
N
R
H
S
H
RNH
2
N N
S
R R
N N
S
R R
H
N
S
R
N
H
R
S
H
S N
S
R
RNH
2
N N
N
R
R R
H
H
2
S
(aqueous)
H
+
+
HS
-
+ H
+
HS
-
+
H
+
+
HS
-
HS
-
HS
-
+
H
+
represents the most acidic proton, in pH >7
this is H
2
O
5, 6
1, 2
3, 4
H
2
S
(gas)
+

The most noticeable physical property of spent triazine solutions is its obnoxious
odor which is primarily due to trace quantities of organosulfur species such as
mercaptans. However, the formation of dithiazine also generates potential problems for
operators in the field, especially its propensity to form solids in aqueous solutions. In
field operations there have been countless instances of dithiazine depositing onto walls of
contactors, pipelines equipped with inline injectors, dump-lines, spent storage tanks and
waste disposal trucks. This deposited material can lead to lower scavenging efficiencies
and in more serious cases to complete blockages. Removal of the material is often costly
in terms of labor, materials and downtime. In general, dithiazine can deposit as either
crystalline or amorphous solids.

Crystalline solids are extremely hard materials that have the appearance of clear,
colourless crystals similar to those observed for large sugar crystals. This type of
ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

7
deposition is rare relative to the formation of amorphous solids and normally forms
slowly at low temperatures (T < 0C) from spent solutions which have low methanol
concentrations. Removal of deposited crystalline dithiazine is extremely problematic.
Melting material out of a system is possible in some cases since the melting points of
crystalline 5 and 6 have been measured at T = 39C
5
and 69C
6
, respectively; however,
care should be exercised with this type of removal since the melted material will re-
solidify upon cooling. Additions of suitable preheated solvents, such as methanol, have
been used to help prevent the re-solidification of the liquid dithiazine. Unfortunately,
heating a system to remove deposits is not always possible and deposits often need to be
removed by time consuming mechanical methods i.e., by hand. Picture #1 (next page)
illustrates a 2 dump-line which contains crystalline dithiazine (compound 5) which had
formed over a three week spring period after the partially spent scavenger had been left in
the system. The dump-line and contactor was completely plugged and later needed to be
replaced with new equipment. This could have been avoided if the system, contactor and
dump- lines, had been completely emptied of all partially spent material.

The majority of dithiazine solids can be classified as amorphous solids which
have a softer, often mushy consistency relative to the harder crystalline solids. However,
these amorphous solids will harden and become brittle if allowed to dry. These materials
are often white or grayish in colour but other colorations have been observed depending
on what other compounds are present. In numerous cases, dithiazine solids also contain
other cyclic organosulfur compounds such as 1,2,4-trithiolane (7) and 1,2,4,6-
tetrathiepane (8) both of which form hard whitish-grey granules that are highly insoluble
in water, alcohols, hydrocarbons and aromatic solvents. Picture #2 illustrates a batch
contactor containing amorphous dithiazine deposits (compound 5) which formed after a
single scavenging run.

S S
S
S S
S S
7 8
1,2,4-Trithiolane 1,2,4,6-Tetrathiepane

ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

8
Picture #1 Crystalline dithiazine (5) deposited in a dump-line



Picture #2 Amorphous dithiazine (5) deposited in a contactor




There are numerous misconceptions about how and why dithiazine solids form;
this section will attempt to simplify the rather complex issue into its basic components.
Initial phases of scavenging lead to low concentrations of 5 or 6 which are completely
soluble. However, as the dithiazine concentrations increase, the dithiazine reaches a
ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

9
saturation point and the solution often becomes turbid. As scavenging continues the
solution will separate into two liquid phases: the top liquid will contain water, amine,
triazine, methanol and small concentrations of dithiazine while the bottom layer will
consist primarily of dithiazine and methanol with a small concentration of water.
Dithiazine forms solids from solutions which become oversaturated with dithiazine due
to changes in temperature (cooling) and as a result of methanol loss in the spent solution.
The kinetics for the formation of solids may not have been studied; however, qualitative
observations suggest that temperature, availability of nucleation sites and agitation are
important factors.

It is often thought that 5 is soluble in water due to its hydroxyl group (HO-);
unfortunately, this reasoning is wrong, for if that were the case, then deposition would
not be a problem with aqueous triazine formulations. In fact, dithiazine 5 has a
calculated solubility of 0.46 g/L in aqueous systems with pH > 7.
7
This value equates to
a solubility of 0.046% (by weight) which is far below spent commercial solutions which
can produce dithiazine concentrations in the range of 30 to 65% (by wt.). The solubility
of 6 in water has been observed by Bakke and Buhaug to be lower than 5; however,
specific solubility data were not reported.
2


The methanol concentration is primarily responsible for determining the
saturation point of dithiazine in spent solutions. In some cases, formulations may not
have a large enough initial concentration of methanol for the maximum quantity of
dithiazine the formulation can generate. In these cases, running the scavenger to high
efficiencies will produce a solution that is supersaturated with dithiazine, a situation
referred to as overspending the chemical. Factors such as large gas flows or higher
contactor temperatures can strip existing methanol out of solution thus producing over
saturated solutions. Another factor which can cause problems is mixing a supersaturated
solution with a poor dithiazine solvent such as water, an operation which has caused
numerous problems in spent chemical tanks and waste disposal trucks.

In order to help mitigate the deposition of dithiazine solids, methanol is often
added to triazine formulations during or after the blending process. Formulations
blended using aqueous formalin solutions will often contain methanol since it is used to
stabilize the formaldehyde in solution. Formaldehyde in water forms long chains (i.e.,
oligomers) which will precipitate out of solution after the chain lengths become too long.
Methanol acts to reduce the size of the oligomers which are more soluble in water by
effectively capping the reactive ends (hydroxyl groups which are hemiacetals) to form
less reactive acetals (i.e., HO-CH
2
-(O-CH
2
)
n
-OH MeO-CH
2
-(O-CH
2
)
n
-OMe).
8,9
In
addition to methanol which is present in commercial formulations, many operators who
have experienced deposition problems will supplement scavenger formulations by adding
methanol directly to the pipelines, contactors, dump-lines or spent chemical storage
tanks.

In general, the methanol concentration should be maintained at the specifications
outlined by the supplier in order to avoid deposition issues. Typically, methanol
ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

10
concentrations in commercial formulations range from 15-35% (by wt.), depending on
the formulation.

Deposition of amorphous dithiazine is often removed by mechanical means such
as pigging pipelines, power spraying solvents for certain contactors and tanks or simply
removal by hand. If solvents are utilized then hot solvents will increase the effectiveness
of the cleaning method, assuming safe application is possible. Many instances exist
where methanol soakings have been used successfully for packed contactors, dump-lines,
spent chemical storage tanks and waste disposal trucks. Unfortunately, there is no
magic bullet approach for the removal of deposited dithiazine and its removal remains
a time consuming and expensive process.

Reaction with Mercaptans

Mercaptans (or thiols) are organo-sulfur compounds which contain a carbon-
bonded sulfhydryl group (R-SH; R represents any carbon moiety) and are the sulfur
analogues of alcohols. Despite what is known regarding the reaction of triazines with
hydrogen sulfide, there is no specific information in the open literature regarding the
reaction of triazines with mercaptans. The current results of a systematic study for the
reaction of two triazines, 1 and 2, with different mercaptans, methanethiol, ethanethiol
and n-propanethiol, will be discussed in subsequent sections.


Results and Discussion

Solutions of 1 and 2 were prepared by mixing molar equivalents of either
monoethanolamine (99.5%) or methylamine (40% by wt.) with formaldehyde (37% by
wt.). The two solutions were characterized using
1
H and
13
C NMR analysis and found to
contain either 51.06% (by wt.) of 1 or 27.08% (by wt.) of 2 and water.

Initial experiments consisted of bubbling helium containing 39.6 1.4 ppm
hydrogen sulfide, 27.1 1.6 ppm methanethiol, 26.3 1.8 ppm ethanethiol and
23.3 1.2 ppm n-propanethiol at a rate of 10 mL/min through a solution of either 1 or 2.
The outlet gas was sampled every twenty minutes for the first 15 hours and then every
three hours until the experiment was complete and sampled gas was analyzed using a gas
chromatograph equipped with a calibrated PFPD.

The uptake experiment for a solution of 1 was determined by measuring the
concentration of hydrogen sulfide and three mercaptans in the outlet gas as a function of
time; the results are illustrated in Figure #1. This experiment was repeated with a
solution of 1 which was diluted with 50% water (by vol.) in order to observe any
dependence on the triazine concentration; the results are illustrated in Figure #2. Each
solution was analyzed via GC/MS at the end of each experiment. Unfortunately, only the
initial triazine, thermal decomposition products and dithiazine were positively identified;
other signals were detected but characterizations were not possible due to insufficient
signal-to-noise ratios.
ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

11
0 10 20 30 40 50 60 70 80 90
0
10
20
30
Time (hrs)
O
u
t
l
e
t

G
a
s

C
o
n
c
e
n
t
r
a
t
i
o
n

o
f

R
S
H

(
p
p
m
)

Figure #1 Concentration of H
2
S (), MeSH (), EtSH () and n-PrSH () in the outlet gas for a solution
of 1.
0 10 20 30 40 50 60 70 80 90
0
10
20
30
Time (hrs)
O
u
t
l
e
t

G
a
s

C
o
n
c
e
n
t
r
a
t
i
o
n

o
f

R
S
H

(
p
p
m
)

Figure #2 Concentration of H
2
S (), MeSH (), EtSH () and n-PrSH () in the outlet gas for a diluted
solution of 1.
ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

12
Initially, the undiluted solution of 1 scavenged 92% of the hydrogen sulfide from
the inlet gas but increased to 100% within two hours of testing. The hydrogen sulfide
uptake remained at 100% over the course of the experiment which was conducted over a
90 hour period. The increased uptake was attributed to an initial increase in temperature
due to the exothermic reaction of 1 with the sulfur species. Higher temperatures will
increase the chemical kinetics of the reaction and will decrease the viscosity of the
solution thus improving gas dispersion (mass transfer). It should be noted, that our test
apparatus is a straight column equipped with stir bar but there is no packing or baffles;
therefore, this phenomenon may not be observed in field units which will, if properly
designed, have better gas-liquid contact than our laboratory set-up. In contrast, the
diluted solution scavenged 100% of the hydrogen sulfide over the duration of the entire
experiment. This improved uptake efficiency during the initial 2 hour test period was
consistent with a lower solution viscosity i.e., better gas dispersion.

In contrast, the concentration of each mercaptan in the outlet gas increased
rapidly after initiation of the gas flow (Figures #1 and #2). The undiluted solution of 1
initially scavenged 89%, 83% and 71% of methanethiol, ethanethiol and n-propanethiol,
respectively but decreased to ca. 3 % for methanethiol and to zero for the other
mercaptans within 50 hours on-stream. Similarly, the diluted solution of 1 initially
scavenged 93%, 91% and 86% of methanethiol, ethanethiol and n-propanethiol,
respectively. The scavenging efficiency then decreased over the course of the experiment
to 2% for methanethiol and to zero for ethanethiol and n-propanethiol.

The plots for the mercaptan concentration in the outlet gases were analyzed using
non-linear least squares regression with an expression for first-order exponential growth
(Y = Y
MAX
(1-e
-kx
)+span). The k terms were calculated to be 0.04164 (R
2
= 0.9992),
0.08045 (R
2
= 0.9984) and 0.0808 (R
2
= 0.9961) for the undiluted solution of 1, while the
k terms for the diluted solution of 1 were determined to be 0.05069 (R
2
= 0.9950), 0.1258
(R
2
= 0.9894) and 0.1712 (R
2
= 0.9863) for methanethiol, ethanethiol and n-propanethiol,
respectively. Larger values for the k terms indicate that there was a faster decrease in the
mercaptan scavenging efficiency i.e., a greater concentration of mercaptan was detected
in the outlet gas within a given time period. It is essential to note, that the k term is not
an absolute rate constant for the chemical reaction of a mercaptan with 1 since neither the
formation of products nor the consumption of reactants were observed directly. In fact,
the k terms may be influenced by a combination of both physical absorption phenomenon
and chemical kinetics.

In practical terms, the efficiency of scavenging mercaptans by a solution of 1 was
found to be inversely related to the molecular weight of the mercaptan; in other words,
smaller mercaptans were more efficiently scavenged from the gas stream than larger
mercaptans. The k terms indicated that the change in scavenging efficiency of
methanethiol was slower relative to those measured for either ethanethiol or n-
propanethiol which had similar changes in their efficiencies.

The 50% (by vol.) dilution of solution 1 decreased the concentration of the
triazine by 50% and subsequently the rate of reaction between the mercaptan and the
ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

13
triazine should have been reduced by 50%. Interestingly, a reduction in the scavenging
efficiency was not observed; in fact, the initial mercaptan scavenging efficiencies
increased after dilution. This result indicated that under our experimental conditions
there are, at least in part, some mass transfer limitations. Thus, the increased uptake
efficiency can be attributed to increasing the gas-to-liquid dispersion due to the reduction
in the viscosity of the solution via dilution with water.

The uptake experiment for a solution of 2 was determined by measuring the
concentration of hydrogen sulfide and three mercaptans in the outlet gas as a function of
time; the results are illustrated in Figure #3. Each solution was analyzed via GC/MS at
the end of each experiment. As was observed for solutions of 1, only the initial triazine
and dithiazine were positively identified; other signals were detected but
characterizations were not possible due to insufficient signal-to-noise ratios.

0 10 20 30 40 50 60 70 80 90
0
10
20
30
Time (hrs)
O
u
t
l
e
t

G
a
s

C
o
n
c
e
n
t
r
a
t
i
o
n

o
f

R
S
H

(
p
p
m
)


Figure #3 Concentration of H
2
S (), MeSH (), EtSH () and n-PrSH () in the outlet gas for a solution
of 2.

The results indicated that the solution of 2 scavenged 100% of the hydrogen
sulfide from the inlet gas over the entire course of the experiment (93 hour duration).
Unfortunately, this result precluded any direct quantitative comparison between the
scavenging rates of hydrogen sulfide for 1 and 2. Qualitatively, it can be stated that
under our test conditions the triazine solutions cannot be distinguished from each other in
terms of their hydrogen sulfide scavenging efficiencies.

ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

14
In contrast, scavenging mercaptans from the inlet gas was determined to be much
less efficient than scavenging hydrogen sulfide. The concentration of each mercaptan in
the outlet gas increased as a function of time (Figures #3). The solution of 2 initially
scavenged 95%, 94% and 89% of methanethiol, ethanethiol and the n-propanethiol,
respectively. The scavenging efficiency decreased to ~9% for methanethiol and to zero,
within the experimental limitations, for the other two examined mercaptans. The plots
for the mercaptan concentration in the outlet gas were analyzed using non-linear least
squares regression with an expression for first-order exponential growth
(Y = Y
MAX
(1-e
-kx
)+span). The k terms were calculated to be 0.03777 (R
2
= 0.9912),
0.06651 (R
2
= 0.9956) and 0.09670 (R
2
= 0.9859) for methanethiol, ethanethiol and n-
propanethiol, respectively. Similar to the argument put forth for the experiments
conducted with solutions of 1, the k terms were not determined by direct measurements
of the product concentration in the solution as a function of time; thus, increased
mercaptan concentrations observed for solutions of 2 were not direct measurement of the
reaction kinetics.

A quantitative comparison could not be made between the mercaptan and
hydrogen sulfide removal since hydrogen sulfide was never detected in the outlet gas
over the course of the experiment. However, the efficiency of scavenging mercaptans by
a solution of 2 was found to be inversely related to the molecular weight of the
mercaptan. The k terms indicated that the change in the scavenging efficiency of
methanethiol was slower than ethanethiol which was slower than n-propanethiol.
Furthermore, the k terms indicated that uptake efficiencies decreased more slowly for the
solution of 2 than for solutions of 1. This observation suggests that the solution 2 would
produce better results for mercaptan scavenging than 1, possible reasons for this result
being elaborated in subsequent sections.

Despite the fact that specific mechanistic information on the chemical reactions
cannot be deduced unequivocally, it is clear that the uptake of mercaptans is rate limited
in comparison to uptake of hydrogen sulfide. Thus, if triazine solutions are utilized in the
field for mercaptan scavenging then an operator should expect a lower scavenging
efficiency and a significantly faster breakthrough relative to scavenging hydrogen sulfide.
The difference in scavenging efficiencies suggests that hydrogen sulfide may be removed
faster at the bottom of a contactor while mercaptans are removed higher in the contactor.
If this is the case, than mercaptan scavenging would work well until the solution becomes
partially spent due to the hydrogen sulfide. One possible method for assuring efficient
mercaptan scavenging would be to introduce fresh scavenger continuously into the
contactor; in other words, run the scavenging system with a flooded tower configuration
rather than a simple batch tower. Highest mercaptan efficiencies may be achieved by
injecting the chemical at the top of the contactor in order to constantly expose the
mercaptans to fresh chemical; in contrast, the hydrogen sulfide would be removed in the
lower portions of the contactor by the partially spent material.

The mechanism for the reaction of mercaptans with triazines has not been
discussed in the open literature and may not have been previously investigated. As an
initial starting point, it is reasonable to assume that the mechanism for the reaction of
ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

15
mercaptans with triazines is similar to the mechanism for reaction of hydrogen sulfide
with triazines. This initial assumption has been used successfully for comparing
nucleophilic substitution reactions of the oxygen analogs of mercaptans i.e., hydroxide
versus methoxide reactivity (HO
-
versus MeO
-
) which differ from each other in terms of
the acidity and nucleophilicity. Similarly, the chemical properties of mercaptans such as
acidity and nucleophilicity are slightly different than hydrogen sulfide due to the carbon
moiety (R-SH). Mercaptans are less acidic than hydrogen sulfide; in fact, methanethiol
has a pKa of 10.3
10
while the pKa of hydrogen sulfide is 7.0
11
i.e., hydrogen sulfide is
almost 2000 times more acidic than methanethiol. The conjugate bases of mercaptans
(RS
-
) are stronger nucleophiles relative to hydrosulfide (HS
-
) which is the conjugate base
of hydrogen sulfide.
12


The first step was to examine how the differences in pKa between the hydrogen
sulfide and mercaptans affect their relative uptake in our experimental set-up. To this
end, an experiment was conducted in which helium containing hydrogen sulfide,
methanethiol, ethanethiol and n-propanethiol was bubbled at a rate of 10 mL/min through
a buffered (pH = 10.0) solution of water i.e., no triazine. The outlet gas was sampled
continuously every twenty minutes for seven hours and was analyzed using a gas
chromatograph equipped with a calibrated PFPD. The solution was buffered at
pH = 10 in order to approximate the solutions of 1 and 2 which had measured pH values
of 10.3 and 9.7, respectively. The results for this experiment are illustrated in Figure #4.

0 1 2 3 4 5 6 7
0
10
20
30
Time (hrs)
O
u
t
l
e
t

G
a
s

C
o
n
c
e
n
t
r
a
t
i
o
n

o
f

R
S
H

(
p
p
m
)


Figure #4 Concentration of H
2
S (), MeSH (), EtSH () and n-PrSH () in the outlet gas for a buffered
(pH = 10) solution of water.

ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

16
The concentration of hydrogen sulfide in the outlet gas steadily increased over the
course of the experiment, a consequence of the acid-base reaction (equation 4) in which
hydrogen sulfide produced hydrosulfide (HS
-
). The increase of hydrogen sulfide in the
outlet gas was attributed to the solution becoming saturated with hydrosulfide; simply
stated, the buffered (pH = 10) aqueous solution functioned as a caustic scrubber for
hydrogen sulfide. Interestingly, the concentration of the mercaptans in the outlet gas
reached inlet gas concentrations within the first twenty minutes. This result indicates that
the dissociation of the mercaptans reaches a rapid equilibrium which favors the
mercaptan and not the thiolate ion (equation 5) in a pH = 10 aqueous solution.



(4)
H
2
S HS
-
+ H
+
S
2-
+ H
+
(5)
RSH RS
-
+ H
+





Changes in the pH of the solution will have two major opposing effects on the
reaction rate since both the equilibrium between the protonated and un-protonated
triazine and the equilibrium between hydrogen sulfide and hydrosulfide will be affected.
For example, decreasing the pH (i.e., making the solution more acidic) will increase the
concentration of the protonated triazine which is a stronger electrophile relative to the un-
protonated triazine thus the rate of reaction will increase. However, the decreased pH
will also shift the equilibrium towards the hydrogen sulfide from hydrosulfide species
which will decrease the rate of reaction since the concentration of the strongest
nucleophile will decrease. In terms of the reaction for hydrogen sulfide, the initial
equilibrium (triazine protonation) predominates since Bakke and co-workers reported that
the overall reaction rate increased in lower pH solutions.
2


The reaction of mercaptans with triazines should be more dependent on the pH of
the solution due to their significantly higher pKa values relative to hydrogen sulfide; this
suggestion assumes that a similar reaction mechanism exists between the reaction of
mercaptans and hydrogen sulfide with triazines. If this assumption is correct, then the
addition of an acid such as hydrogen sulfide to either a solution of 1 or 2 should produce
more efficient mercaptan uptake. In fact, a control uptake experiment for a solution of 1
was conducted by measuring the concentration of the three mercaptans in the outlet gas
as a function of time in the absence of hydrogen sulfide. The results for this experiment
are illustrated in Figure #5.



ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

17
0 10 20 30 40 50 60
0
10
20
30
Time (hrs)
O
u
t
l
e
t

G
a
s

C
o
n
c
e
n
t
r
a
t
i
o
n

o
f

R
S
H

(
p
p
m
)

Figure #5 Concentration of MeSH (), EtSH () and n-PrSH () in the outlet gas for a solution of 1
with no H
2
S in the feed gas.

The solution of 1 in the absence of hydrogen sulfide initially scavenged 85%,
76% and 68% of methanethiol, ethanethiol and the n-propanethiol, respectively. The
concentration of each mercaptan in the outlet gas increased as a function of time. In fact,
the scavenging efficiency for each mercaptan decreased to ~4% for methanethiol and to
zero, within the experimental limitations, for the other two examined mercaptans.

The results indicate that the absence of hydrogen sulfide did decrease the
mercaptan scavenging efficiency. This result is consistent with our initial hypothesis that
the initial steps in mercaptan scavenging occurs due to nucleophilic attack of the
protonated triazine by the thiolate ion (Scheme #3; top reaction pathway involving the
triazine) which is positively influenced by the presence of a weak acid such as hydrogen
sulfide. It should be noted, that the hydrogen sulfide used in previous experiments were
done using very low concentrations (<40 ppm); thus more significant scavenging
efficiencies would be expected with higher concentrations of hydrogen sulfide. Although
the current studies are consistent with the mechanism illustrated in Scheme #3, these are
preliminary studies and greater detailed mechanistic work is required to elucidate aspects
of this complex reaction system.

The results indicated that the reactions of mercaptans with triazines such as 1 are
affected by the pH of the solution. This result suggest that the minor difference in
efficiency observed for scavenging mercaptans between solutions containing 2 and 1
could be attributed to the pH of the solution and not due to the structure of the triazine.
ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

18
In other words, scavenging mercaptans with more acidic triazine solutions will produce
higher scavenging efficiencies; however, this strategy has its limits because if the
solution becomes too acidic then hydrolysis reactions will render the scavenger solution
inert.


Scheme 3

N N
N
R
R R
H
N N
R R
N
R
H
S
R
1
N N
N
R
R R
H
N N
N
R
R R
N N
N
R
R R
N N
R R
N
R
S
R
1
N N
R R
N
R
H
S
R
1
R
1
SH
(aqueous
R
1
S
-
+ H
+
R
1
SH
(gas)
H
+
+
R
1
S
-
R
1
S
-
+
R
1
S
-
R
1
S
-
H
+
-


Alternatively, an increase in pH may also result in an increased mercaptan
scavenging efficiency since at a higher pH the concentration of the thiolate ion would
increase. Although there would be a lower concentration of protonated triazine in more
basic solutions, the thiolate ion is a much stronger nucleophile than either hydrosulfide or
its conjugate acid (RSH) and may react rapidly with the un-protonated triazine (Scheme
#3; bottom reaction pathway involving the triazine). This possibility was tested using an
experiment in which helium containing methanethiol, ethanethiol and n-propanethiol was
bubbled at a rate of 10 mL/min through a solution of 1 (3 mL) which had its pH adjusted
to pH = 14 by adding 0.2814 grams (0.33 M) of potassium hydroxide to 15 mL of a
solution of 1. A control experiment also was conducted using an aqueous solution of
(0.33 M) potassium hydroxide. The results for these experiments are illustrated in Figure
#6a,b.

The solution of 1 with added 0.33 M KOH initially scavenged 93%, 93% and
86% of methanethiol, ethanethiol and the n-propanethiol, respectively and after 93 hours
decreased to 34%, 22% and 22% for each respective mercaptan. Similarly, the aqueous
0.33 M KOH solution initially scavenged 94%, 95% and 89% of methanethiol,
ethanethiol and the n-propanethiol, respectively: this decreased after 93 hours to 21%,
ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

19
11% and 3% for each respective mercaptan. Thus, the difference in scavenging
efficiencies after 93 hours between the triazine and control experiment were 13%, 11%
and 19% for methanethiol, ethanethiol and n-propanethiol, respectively. It should be
noted, that the solution of 1 with added KOH (Figure #6a; pH = 14) was significantly
more efficient at scavenging mercaptans than the neat solution of 1 (Figure #5; pH =
10.3). In each case, the concentrations of mercaptans increased overtime. The growth
rate in the concentration of each mercaptan were slower for the solution of 1 (pH = 14)
relative to the aqueous KOH solution.


0 10 20 30 40 50 60 70 80 90
0
10
20
30
Time (hrs)
O
u
t
l
e
t

G
a
s

C
o
n
c
e
n
t
r
a
t
i
o
n

o
f

R
S
H

(
p
p
m
)
0 10 20 30 40 50 60 70 80 90
0
10
20
30
Time (hrs)
O
u
t
l
e
t

G
a
s

C
o
n
c
e
n
t
r
a
t
i
o
n

o
f

R
S
H

(
p
p
m
)


Figure #6a,b (left) Concentration of MeSH (), EtSH () and n-PrSH () in the outlet gas for a solution
of 1 (pH = 14); (right) concentration of MeSH (), EtSH () and n-PrSH () in the outlet
gas for an aqueous solution of pH = 14.


The results indicated that triazine solutions treated with a strong base (pH > 13.8)
were more effective at scavenging mercaptans than the untreated triazine solutions
(pH = 8-11). The observed increase in scavenging efficiency was attributed primarily to
scavenging by the caustic material (i.e., KOH); however, the affect on the triazine
chemistry was noticeable in the latter parts of the experiment. Although Bakke and co-
workers assumed that only the protonated triazine could react with the hydrosulfide in
solution
2
, the current results suggest the thiolate ions which are much stronger
nucleophiles than hydrosulfide do react with the un-protonated triazine and this pathway
becomes more dominant in highly basic solutions (Scheme #3). Caution should be
exercised in adding caustics to triazine solutions for field applications due to potential
safety issues and the potential for the formation of scale.





ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

20
Conclusion

Triazine based scavengers have been found to efficiently remove low levels of
hydrogen sulfide from gas streams. This reaction produces dithiazines which under some
conditions can form solid deposits. Despite the erroneous concept that dithiazines (or at
least some dithiazines) are water soluble, deposition of dithiazine occurs primarily in
systems which have insufficient methanol concentrations. Mitigation of dithiazine
deposition can be accomplished by supplementing formulations with methanol or by
utilizing the scavenger at lower efficiencies.

It has been suggested that triazines can be used to remove mercaptans effectively
from hydrocarbon streams. The current investigation shows that scavenging of
mercaptans was less efficient than for hydrogen sulfide using fresh triazine and the
efficiency of their removal fell exponentially over the course of scavenging experiments
whereas 100% removal of hydrogen sulfide was maintained. It has been suggested that
lower observed efficiencies were due to the lower acidity (larger pKa values) of the
mercaptans relative to hydrogen sulfide. The current work also suggests that thiolate
ions, in contrast to hydrosulfide, can react with the un-protonated triazine in solution and
that this mechanistic pathway becomes more dominant in more basic solutions. Future
work from our laboratory will explore the reaction in greater detail and possible ways of
increasing the practical usefulness of triazines for mercaptan scavenging.


Experimental

Low resolution mass spectra were recorded using a Varian 450 gas
chromatograph equipped with a Varian 240-MS detector and a VF-5ms fused silica
capillary column (Varian, Inc.).
1
H and
13
C NMR were recorded on a Bruker DRX-400
spectrometer in D
2
O or MeOD solution and were referenced using the solvent. The
spectra for the reaction between gaseous mercaptans and hydrogen sulfide with 1 and 2
were obtained using a Varian 3800 GC equipped with a calibrated PFPD detector and a
RT-U-bond column (Restek).

Methyl amine (Aldrich), monoethanolamine (Aldrich), 37% (wt.) aqueous
formaldehyde solution (Aldrich), D
2
O (Aldrich) and potassium hydroxide (J.T.Baker)
were the highest purity available and were used as received.

1,3,5-Tris(2-hydroxyethyl)hexahydro-s-triazine (1) was synthesized according to
the following procedure. An aqueous solution of 37% (by wt.) formaldehyde (103.05 g;
1.27 moles) was added dropwise over a 1 hour period to monoethanolamine (78.22 g;
1.27 moles) in a 500 mL round bottom. The solution was stirred for a 6 hour period. The
solution had a measured pH = 10.26. The diluted triazine formulation was produced by
adding 50% (by vol) RO water; this solution had a measured pH = 10.27. The product
was identified on the basis of the following spectroscopic data:
1
H NMR (D
2
O), = 2.71
(t, 6H), 3.38 (s, 6H), 3.72 (t, 6H);
13
C NMR, = 53.69, 58.72, 72.89.

ASRL Quarterly Bulletin No.155 Vol. XLVII No. 3, October-December 2010, pp. 1-21.

21
1,3,5-Trimethylhexahydro-s-triazine (2) was synthesized according to the
following procedure. An aqueous solution of 37% (by wt.) formaldehyde (144.20 g; 1.78
moles) was added dropwise over a 1 hour period to 40% (by wt.) methylamine (137.95 g;
1.78 moles) in a 500 mL round bottom. The solution was stirred for a 6 hour period.
The solution had a measured pH = 9.69. The product was identified on the basis of the
following spectroscopic data:
1
H NMR (D
2
O), = 2.31 (s, 9H), 3.39 (s, 6H);
13
C NMR,
= 39.01, 74.70.

Reaction of mercaptans and hydrogen sulfide with the triazines were studied
using the following procedure. A sample of the triazine solution (3.0 mL) was placed in
a 10 mL graduated cylinder equipped with a stir bar. The solution was stirred while an
inlet gas stream was bubbled through the solution at a rate of 10 mL/min. The inlet gas
was composed of 39.6 1.4 ppm H
2
S, 27.1 0.8 ppm MeSH, 26.3 0.9 ppm EtSH,
23.3 0.6 ppm n-PrSH and balanced with helium. The mercaptans were supplied via
permeation tubes heated in a GC oven at a constant temperature of T = 50C. The gas
was collected from the capped graduated cylinder via a 1/16 tubing into a gas injection
loop on a Varian 3800 gas chromatograph equipped with a calibrated PFPD detector.
The gas chromatograph was automated to sample continuously every 20 min. Reaction
solutions were analyzed using GC/MS.


Acknowledgements

We wish to thank Kevin L. Lesage for his technical assistance. The financial
support of the membership of Alberta Sulphur Research Ltd. is also greatly
acknowledged.

References

1) Bakke, Jan M.; Buhaug, Janne B.; Riha, J. Ind. Eng. Chem. Res. 2001, 40,
6051-6054.
2) Bakke, Jan M.; Buhaug, Janne B. Ind. Eng. Chem. Res. 2004, 43, 1962-1965.
3) Pretty, Jack; Glaser, Robert; Jones III, Johny; Lunsford, R. Alan. Analyst 2004, 129,
1150-1155.
4) Taylor, Grahame N.; Matherly, Ron. Ind. Eng. Chem. Res. 2010, 49, 5977-5980.
5) Galvez, Ruiz; Carlos, Juan. Inorganic Chemistry, 2003, 42, 7569-7578.
6) Aleev, R.S. Doklady Akademii Nauk SSSR, 1988, 303, 873-875.
7) Calculated using Advanced Chemistry Development (ACD/Labs) Software V8.14
(1994-2010 ACD/Labs)
8) Grtzner, Thomas; Hasse, Hans. J. Chem. Eng. Data 2004, 49, 642-646.
9) Hahnenstein, I.; Hasse, H.; Kreiter, C. G.; Maurer, G. I & EC Res. 1994, 33, 1022-
1029.
10) Stewart, R. The Proton: Applications to Organic Chemistry, Academic Press: New
York, 1985.
11) Ionization Constants of Organic Acids in Solution, Serjeant E. P. and Dempsey B.,
Eds.; IUPAC Chemical Data Series No. 23, Pergamon Press, Oxford, UK, 1979.
12) Kyllonen, Arja; Koskikallio, Jouko. Suomem Kemistitehti, 1972, 45(7-8), 212-4.

You might also like