Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

REFRACTORY NITROGEN COMPOUNDS IN HYDROCRACKING PRETREATMENT–

IDENTIFICATION AND CHARACTERIZATION

Angelica Hidalgo-Vivas1, Peter Wiwel1, Berit Hinnemann1, Per Zeuthen1,


Bent O. Petersen2 and Jens Ø. Duus2

1
Haldor Topsøe, Nymøllevej 55, 2800 Lyngby, Denmark
2
Carlsberg Laboratory, Gamle Carlsberg Vej 10, 2500 Valby, Denmark

Abstract: Currently, there is a growing need to hydroprocess heavier and tougher crude
oils with increased nitrogen content. Therefore, hydrodenitrogenation (HDN) has become
a critical hydroprocessing reaction, making it essential to gain insight into which
nitrogen-containing compounds are the most difficult to treat. In the present paper, we
describe the identification of nitrogen compounds in severely pretreated feed for
hydrocracking. The nitrogen compounds in the N-slip to the hydrocracker are isolated
and concentrated on solid-phase extraction (SPE) columns and identified by gas
chromatography mass spectrometry (GC-MS), gas chromatography atomic emission
detection (GC-AED) and nuclear magnetic resonance (NMR) spectroscopy. Density
functional theory (DFT) calculations support the structural identification and are further
used to investigate reactivity We find that the most refractory organic nitrogen
compounds in the N-slip belong to the family of 4,8,9,10-tetrahydro-
cyclohepta[def]carbazoles. This is in accordance with previous studies (Zeuthen et al.,
2001) which pointed at alkyl substituted carbazoles as generally being the most refractory
species. Our findings can be related to several studies using model feedstocks which
show that in particular basic nitrogen-containing molecules inhibit HDS reactions the
most. The identified 4,8,9,10-tetrahydro-cyclohepta[def]carbazoles are slightly more
basic than other carbazoles and thus likely to have an impact on the performance of the
downstream catalysts; however, these species show extremely low reactivities and they
are extremely difficult to remove under normal hydrotreating conditions.

Keywords: carbazoles, hydrodenitrogenation, VGO, GC-AED, GC-MS, NMR, DFT

1. INTRODUCTION

The current economic and oil reserves conditions worldwide have resulted in a growing interest in processing heavy
oils and even extra heavy oils with much higher nitrogen content. It is our experience that there has been a 20%
increase in nitrogen content of feeds to hydrocrackers over the past years (Bettati et al., 2005). Removal of nitrogen
is essential to prevent catalyst poisoning in downstream refinery processes, e.g. hydrocracking, catalytic cracking,
and reforming. Organic nitrogen is removed catalytically by hydrodenitrogenation (HDN), which is generally the
most difficult hydrotreatment reaction (Topsøe et al., 1996).

Most of the nitrogen is present as heterocycles with multiple aromatic rings. Basic compounds are mainly 6-member
ring nitrogen compounds, e.g., quinolines, benzoquinolines, etc. Non-basic compounds are mainly 5-member ring
compounds, e.g., indoles, carbazoles, etc. For reference, some chemical structures are shown in Table 1.

Typically, half the total nitrogen is concentrated in the heaviest 30% in heavy feeds (Bettati et al., 2005) with
carbazole compounds compounds substituted at position 1 being the most abundant (Wiwel et al., 2000). The high
relative concentrations of 1-substituted carbazoles in crudes are attributed to preferential migration of nitrogen
compounds having sterically shielded N-centres. Di- and tri-methylcarbazoles with substitution at position 1 were
observed to be the most predominant.
Table 1. Structures of selected organic nitrogen compounds

Basic Non-basic
Quinoline f d Indole 4
3
g c 5

b.p. 238oC b.p. 254oC 2


h b 6 N
N
7 H

Benzo(h)quinoline Carbazole 5 4
6 3

b.p. 338oC b.p. 355oC 7


2
N
8 N 1
H
Acridine 1,2-Benzocarbazole
(dibenzo(b,e)pyridine or
benzo(b)quinoline)
N
b.p. 346oC N
H

Considerable attention has been paid to elucidate the problem of nitrogen compound inhibition because the effects
are not only important from a process point of view, but also influence the development of hydroprocessing
catalysts. Organic nitrogen compounds have a significantly negative kinetic effect on hydrotreating reactions like
hydrodesulphurisation (HDS), on other hydrogenolysis reactions and on hydrogenation reactions. The negative
effect on hydroprocessing catalysts, both hydrotreating catalysts and in particular on the more acidic hydrocracking
catalysts, and consequently on the performance of the hydrocrackers has been accepted for decades (Gary and
Handwerk, 1994).

Many general studies related to this topic have been made with the primary objective of improving the knowledge of
HDS and HDN (Girgis and Gates, 1991; Laredo et al., 2004; Owusu-Boakye et al., 2005). These studies have not
focused on the characterization of the most refractory nitrogen compounds and on the understanding which of the
compounds present in the N-slip are the worst poisons to the HC catalysts or on understanding how this inhibition
occurs. In the area of hydrotreated stream compositions, prior work identified alkylcarbazoles as low reactivity N-
compounds (Dorbon et al., 1984; Igniatiadis at al., 1986), but little information on the relative rates of individual
compound conversions was gained because of the difficulties in quantifying the results.

In this paper, we describe the detection and identification of a group of very refractory nitrogen-containing organic
compounds, present in the 340-370oC distillate fraction that persists even after high severity HDN and describe the
impact on the HC catalyst performance.

2. EXPERIMENTAL AND COMPUTATIONAL METHODS

2.1. VGO feed and hydrotreated gas oil product samples

In this study, a vacuum gas oil and its hydrotreated products were investigated. The feed, with properties shown in
Table 2, was hydrotreated in a pilot plant consisting of one isothermal reactor loaded with a commercial
NiMo/alumina catalyst. The catalyst size used was 1/20” trilobe and the catalyst bed was diluted by 40 vol% inert
carborundum (SiC) prior to loading in order to improve liquid distribution in the reactor. Pure hydrogen was used in
once-through mode. The test conditions at which the pilot plant feed was treated and properties of the products are
presented in Table 3. Characterization and distribution of the nitrogen-containing compounds of four of the products
were also obtained.
Table 2. VGO feed properties

Property Method VGO feed Property Method VGO feed


S, wt% D 4294 2.2960 Simulated Dist. D 2887
N, ppm (w/w) D 4629 1394 0.5 wt% (IBP), °C 301
H, wt% D 4808 12.16 5 wt%, °C 353
SG 60/60°F D 4052 0.9267 10 wt%, °C 372
Flash Point, °C D 93 211 20 wt%, °C 398
Pour Point, °C D 5949 40 30 wt%, °C 417
Viscosity at 50°C, cSt D 445/446 44.1 40 wt%, °C 434
Viscosity at 100°C, cSt D 445/446 8.1 50 wt%, °C 449
Aromatics IP 391 60 wt%, °C 465
1 Ring, wt % 16.2 70 wt%,°C 483
2 Ring, wt % 8.74 80 wt%, °C 502
3+ Ring, wt % 14.89 90 wt%, °C 528
N Basic, wt ppm HTAS 493 95 wt%. °C 546
Fe/Ni/V HTAS <1/ <1/ <1 99.5 wt% (FBP), °C 588

Table 3. Hydrotreated products

Test Conditions Condition 1 Condition 2 Condition 3 Condition 4


o
Temperature C 380.4 380.8 398.8 400.9
Pressure barg 60 120 60 80
LHSV 1/h 0.51 0.51 0.52 0.52
H2/oil Nm3/m3 770 767 764 753
Property Method HDT product 1 HDT product 2 HDT product 3 HDT product 4
S, wt% D 4294 0.0068 0.0035 0.0003 0.0010
N, wt ppm D 4629 162 14 23 7
H, wt% D 4808 13.20 13.58 13.18 13.48
SG 60/60°F D 4052 0.8846 0.8767 0.8697 0.8593
Aromatics IP 391
1 Ring, wt % 29.07 22.3 27.43 22.61
2 Ring, wt % 4.9 1.74 5.54 2.8
3+ Ring, wt % 3.14 0.79 3.48 1.76
Sim. Dist. D 2887
IBP/ 5/ 10wt%, °C 114/251/306 113/246/300 96/169/221 81/156/205
30/ 50/ 70 wt%,°C 381/422/458 379/420/457 334/390/435 317/382/431
90/ 95/ FBP, °C 509/530/576 508/530/576 494/517/570 491/515/567

2.2. Isolation and identification of refractory nitrogen compounds

Prior to analysis, the nitrogen containing compounds in the 340-370oC product cut were up-concentrated by the use
of silica-based solid-phase extraction (500 mg Si-SPE) using a procedure similar to that described by Li et al.
(1997). Non-polar compounds were flushed through the column whilst the polar compounds, including all the
nitrogen-containing compounds, were trapped on the silica SPE-column. These can then be separated into
compound classes and recovered by flushing the SPE column with solvents of increasing strength.

Subsequently, we used sensitive and selective methods and techniques, namely GC-MS, GC-AED and NMR to
obtain quantitative results for the content and identification of these extremely refractory N-compounds in severely
hydrotreated products. Furthermore, we used density functional theory (DFT) calculations to support the structural
identification and to investigate reactivity.
3. RESULTS AND DISCUSSION

Chromatograms of the VGO feed and products hydrotreated at a relatively two temperatures (380 and 400oC) are
shown in Figure 1. Carbazoles and benzocarbazoles appear in the chromatogram of the product with 160 ppm (w/w)
total nitrogen content, whilst only traces of benzocarbazoles appear in the other products. In the chromatograms of
the products treated at a higher temperature (400oC), one main peak and several minor peaks of unidentified
nitrogen-containing compounds dominate. The relative amount of these main unidentified compound increases with
lower total amount of nitrogen, indicating, that this compound, or compound group, must be very unreactive and
difficult to remove.

Figure 1. GC-AED Nitrogen chromatograms of a) feed; b, c) hydrotreated products at 380oC; d, e) hydrotreated


products at 400oC. Other conditions: LHSV=0.5 h-1 and H2/oil= 770 Nm3/m3
3.1 Identification of the unknown family, MW determination and structure

The sample was analysed by GC/MS in an attempt to identify this un-reactive compound. The mass spectrum of the
unknown compound showed that the molecular weight was 207, corresponding to a molecular formula of C15H13N.
Mono-methyl substituted compounds on an identical molecular skeleton were also identified. Several possible
structures were proposed based on the chemical formula, but a positive identification based on mass spectrum alone
can not be done, primarily because isomeric compounds can not be distinguished. To elucidate the molecular
structure and the 3-D spatial orientation of the unknown compounds, NMR and DFT have been employed.

In order to use NMR it has been necessary to purify the compound in sufficient amounts to be able to measure the
spectrum. By repeated SPE purification it was possible to obtain a sample with up to 70% purity of the unknown
compound, the rest being the methyl-substituted compound. NMR spectroscopy has been done on the purified
sample and results confirmed the structure shown in Figure 2 as the correct one among those proposed.
N
H
Figure 2. Proposed structure of the carbazole isomer 4,8,9,10-tetrahydrocyclohepta[def]carbazole
4,8,9,10-Tetrahydro-cyclohepta[def]carbazole

To further supplement the NMR assignment, different structural isomers of the C15H13N molecule have been
calculated by DFT (only two are shown in Figure 4). The 4,8,9,10-tetrahydro-cyclohepta[def]carbazole isomer is
more stable than the next isomer by 16 kJ/mol. This energy difference is sufficient to identify the 4,8,9,10-
tetrahydro-cyclohepta[def]carbazole isomer as the most stable one among those investigated.

Figure 3. Structure of 4,8,9,10-tetrahydrocyclohepta[def] carbazole and 5,6-dihydro-4H pyrido [3,2,1-jk]carbazole.


Colour code for the molecular structures is: white (hydrogen), grey (carbon), blue (nitrogen).

3.2. Hydrogenation reactivity by DFT and hydrogenation equilibrium


Studies of HDN of various N-heterocyclic compounds (Cochetto and Satterfield, 1976; Girgis and Gates, 1991;
Topsøe et al., 1996), including carbazole, have shown that the major pathway involves hydrogenation of the N-ring,
followed by cleavage of the C-N bond, forming an amine intermediate and finally hydrogenolysis of the amine to
hydrocarbons and ammonia.

DFT is used to investigate the differences in hydrogenation energies and it is initially determined that hydrogenation
is energetically less favourable for 4,8,9,10-tetrahydrocyclohepta[def]carbazole than for carbazole itself. This
difference in hydrogenation might be one of the factors responsible for its low reactivity.

4,8,9,10-tetrahydro-cyclohepta[def]carbazoles are scarcely present in the feed; however, they predominate in the
products from hydrotreatment at high temperatures, suggesting that their presence is associated to thermodynamic
equilibrium limitations to the hydrogenation step. Comparison of the chromatograms of the products treated at a
pressure of 60 bar (Figure 1b and Figure 1d) show that other heavy benzocarbazoles do not seem to have reached
equilibrium, as they disappear by an increase in reaction temperature. Figure 1c shows that high hydrogen pressure
is not sufficient to remove them as the pressure was double that of the Figure 1b.

3.4. Impact of nitrogen on HC catalyst performance

The catalysts in downstream processes, e.g., hydrocracking (HC) catalysts, will come into contact with the N-
compounds that survive hydrotreatment, i.e., the identified family of 4,8,9,10-tetrahydrocyclohepta[def]carbazoles
as we have already established their low reactivity and difficulty to remove under normal hydrotreating conditions.

Moreover, based on the affinity to silica, it is possible to establish that the 4,8,9,10-tetrahydro-
cyclohepta[def]carbazoles are slightly more basic than other carbazoles. It is thus expected that the adsorption on the
acid sites of HC catalysts will be stronger than that of NH3, thus inhibiting the cracking activity to a greater extent.
Furthermore, a detrimental effect to the hydrogenation step in the hydrodesulphurisation, hydrocracking and
saturation of other molecules is also expected, as it has been concluded from real feed experiments on VGO that
show that heavy organic nitrogen-containing compounds are the major inhibitors to the hydrogenation function
during HDS (Hidalgo-Vivas et al., 2003).

The consequence of organic nitrogen into the HC catalyst is to reduce conversion and liquid volume yields. The
increase in operating temperature to maintain conversion reduces the cycle length and may affect the products
quality negatively, mainly by increasing products density. Equilibrium limitations in aromatics saturation will be
favoured by increasing temperature, thus affecting the cetane number of the diesel product, smoke point of the jet
fuel and viscosity index of the unconverted oil. The distribution of the products is also affected by the increase in
temperature. The ratio of iso to normal paraffins ratio in the product decreases at higher temperatures because the
cracking rate of isoparaffins is faster than that of linear paraffins. Gas yield increases whilst naphtha and jet yields
suffer. Hydrogen consumption will also increase. The useful life of HC catalysts decreases rather rapidly with
increasing nitrogen content. It may be necessary to reduce throughput or end boiling point of the VGO hydrocracker
feed.

ACKNOWLEDGEMENT

The authors wish to thank Dr. Michael Brorson for his knowledge in order to give a systematic name of the 7-ring
carbazole isomer, and Dr. Kim G. Knudsen for ongoing discussion throughout the identification process.

REFERENCES

Bettati, A., Zeuthen, P.; Hidalgo-Vivas, A (2005) Nitrogen tolerance of hydrocracking catalysts. In Topsøe
Hydrocracking Catalyst and Technology Seminar 2005, Copenhaguen
Cochetto, J.F.; Satterfield, C.N. (1976) Thermodynamic equilibria of selected heterocyclic nitrogen compounds with
their hydrogenated derivatives. Ind. Eng. Chem. Proc. Res. Dev. 15(2), 272-277
Dorbon, M.; Igniatiadis, I.; Schmitter, J.M.; Arpino, P.; Guiochon, G.; Toulhout, H.; Huc, A.(1984) Identification of
carbazoles and benzocarbazoles in a coker gas oil and influence of catalytic hydrotreatment on their distribution.
Fuel 63 (4), 565
Gary, J.H.; Handwerk (1994), G.E. Petroleum Refining. Technology and Economics. 3rd. ed., Marcel Dekker, N.Y.
USA p. 153, 189-193
Girgis, M.J.; Gates, B.C. (1991) Reactivities, Reaction Networks, and Kinetics in High-Pressure Catalytic
Hydroprocessing. Ind. Eng. Chem. Res. 30, 2021-2058
Hidalgo-Vivas, A.; Knudsen, K.G.; Zeuthen, P. (2003) Kinetic modelling of VGO hydrotreatment in FCC
pretreatment service. AIChE Spring Meeting, Paper T8a12a.
Igniatiadis, I.; Kuroki, M.; Arpino, P. (1986) Identification of carbazole derivatives in a hydrotreated coker gas oil
by gas chromatography and gas chromatography - mass spectrometry. J. Chromatogr. 366, 251-260
Laredo, G.C.; Montesinos, A.; De los Reyes, J.A. (2004) Inhibition effects observed between dibenzothiophene and
carbazole during the hydrotreating process. Appl. Catal. A 265 (2), 171-183
Li, M.; Yao, H.; Stasiuk, L.D.; M.G. Fowler, M.G.; Larter, S.R.(1997) Effect of maturity and petroleum expulsion
on pyrrolic nitrogen compound yields and distribution in Duvernay Formation petroleum source rocks in Alberta,
Canada. Organic Geochemistry 26, 731-744.
Owusu-Boakye, A.; Dalai, A.K.; Ferdous, D.; Adjaye, J. (2005) Experimental and Kinetic Studies of Aromatic
Hydrogenation, Hydrodesulfurization, and Hydrodenitrogenation of Light Gas Oils Derived from Athabasca
Bitumen. Ind. Eng. Chem. Res. 44 (21), 7907-8156
Topsøe, H.; Clausen, B.; Massoth, F. (1996) Hydrotreating Catalysts. Science and Technology. Springer-Verlag,
Germany.
Wiwel, P.; Knudsen, K.; Zeuthen, P.; Whitehurst, D. (2000) Assessing compositional changes of nitrogen
compounds during hydrotreating of typical diesel range gas oils. Ind. Eng. Chem. Res. 39(2), 533-540
Zeuthen, P.; Knudsen, K.; Whitehurst, D.D. Organic nitrogen compounds in gas oil blends, their hydrotreated
products and the importance to hydrotreatment. Catal. Today 2001, 65(2-4), 307-314

You might also like