You are on page 1of 43

Author's Accepted Manuscript

Flow characteristics and rheological properties of natural gas hydrate slurry in the presence of anti-agglomerant in a flow loop apparatus Ke-Le Yan, Chang-Yu Sun, Jun Chen, Li-Tao Chen, De-Ji Shen, Bei Liu, Meng-Lei Jia, Meng Niu, Yi-Ning Lv, Nan Li, Zhi-Yu Song, Shu-Shan Niu, Guang-Jin Chen
www.elsevier.com/locate/ces

PII: DOI: Reference:


To appear in:

S0009-2509(13)00750-1 http://dx.doi.org/10.1016/j.ces.2013.11.015 CES11395


Chemical Engineering Science

Received date: 28 July 2013 Revised date: 22 October 2013 Accepted date: 10 November 2013 Cite this article as: Ke-Le Yan, Chang-Yu Sun, Jun Chen, Li-Tao Chen, De-Ji Shen, Bei Liu, Meng-Lei Jia, Meng Niu, Yi-Ning Lv, Nan Li, Zhi-Yu Song, ShuShan Niu, Guang-Jin Chen, Flow characteristics and rheological properties of natural gas hydrate slurry in the presence of anti-agglomerant in a flow loop apparatus, Chemical Engineering Science, http://dx.doi.org/10.1016/j.ces.2013.11.015 This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting galley proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Flow characteristics and rheological properties of natural gas hydrate slurry in the presence of anti-agglomerant in a flow loop apparatus
Ke-Le Yana, Chang-Yu Suna,, Jun Chena, Li-Tao Chenb, De-Ji Shena, Bei Liua, Meng-Lei Jiaa, Meng Niua, Yi-Ning Lva, Nan Lia, Zhi-Yu Songa, Shu-Shan Niua, Guang-Jin Chena,* a. State Key Laboratory of Heavy Oil Processing, China University of Petroleum, Beijing 102249, China b. Center for Hydrate Research, Chemical & Biological Engineering Department, Colorado School of Mines, Golden, Colorado, United States

ABSTRACT: The flow characteristics and rheological properties of natural gas


hydrate slurry, with initial water cuts ranging from 5 to 30 vol%, were investigated in a flow loop. The experimental results indicate that the hydrate slurry can be considered a pseudoplastic fluid and presents more obvious shear-thinning behaviour with the increase in the hydrate volume fraction. The study on the fluid morphology demonstrated that the original structure of the water-in-oil emulsion is destroyed by the formation of gas hydrate, and the hydrate slurry is ultimately transported as a solid dispersion system. An empirical Herschel-Bulkley-type equation that considers the hydrate volume fraction was developed to improve the description of the rheological properties of the hydrate slurry. The apparent viscosities of the hydrate slurry calculated by the new equation were in accordance with the experimental data. Shutting-down/restarting tests using three shutting-down times (2 h, 4 h, and 8 h) were performed.The experimental results indicate that the hydrate slurry can be easily and safely restarted from the static state after a long shutting-down period and exhibits obvious thixotropic behaviour with increasing shutting-down time. Keywords: hydrate; slurry; rheological properties; flow characteristics;

anti-agglomerant

To whom correspondence should be addressed. Fax: +86 10 89733156. E-mail: cysun@cup.edu.cn (C. Y. Sun), gjchen@cup.edu.cn (G. J. Chen). 1

1. Introduction
Gas hydrates are ice-like clathrate-type crystals in which cages of water molecules are stabilised by the host molecules (Sloan and Koh, 2008). During oil/gas exploitation, light alkanes, such as methane, ethane, and propane, can form gas hydrates with the water produced in pipelines under high pressure and relatively low temperature. The hydrate particles can agglomerate into hydrate plugs, which may cause total blockage (Gao, 2009). Recently, the problem caused by natural gas hydrate blocks has become increasingly severe with the increasing water depth of the offshore oil and gas pipelines. According to a survey, the annual cost of preventing the hydrate issue is over U.S. $200 million (Sloan, 2003) and accounts for 5 to 8% of the total product plant cost (Sloan et al., 2011; Chandragupthan, 2011). Many approaches are used to prevent hydrate plugs (Kelland, 2006). The most commonly method is the addition of thermodynamic inhibitors, e.g., methanol or ethylene glycol, which prevent hydrate formation by shifting the hydrate equilibrium curve toward higher pressures and lower temperatures to keep the operation conditions outside the hydrate stability region. However, the concentration usually required for these inhibitors to be effective is 30 to 50 wt% of the water mass. Such a high concentration requires a large amount of additives to be used, increasing the cost of project operation and requiring the reprocessing of wastewater. Low-dosage hydrate inhibitors (LDHIs), including kinetic hydrate inhibitors (KHIs) and anti-agglomerants (AAs), have been researched and developed for many years as an alternative method to control gas hydrates (Kelland, 2006; Arjmandi et al., 2005).
2

KHIs are a type of water-soluble polymer with functional groups that can be accommodated into clathrate hydrate cages. Unlike thermodynamic inhibitors, they can delay hydrate nucleation (usually crystal growth as well), providing sufficient time to transport the fluids to the process facilities before hydrate plugs build up in the pipeline. KHIs have been widely applied for hydrate inhibition in gas-dominated systems in Qatar and Iran, where the subcooling is low (<10 K) and the subcooling maximum is seasonal (Creek, 2012). However, it is well-known that they may cease to function when the subcooling is over approximately 10 K (Mehta et al., 2003). AA molecules generally have a hydrophilic head and a hydrophobic tail. They allow gas hydrates to form but keep the hydrate crystals small and dispersed in the hydrocarbon phase. The hydrate crystals can then be transported as non-sticky hydrate slurry in the pipelines (Kelland, 2006; Huo et al., 2001). Both KHIs and AAs are added at relatively low concentrations. Approximately 0.1 to 3.0 wt% of the aqueous phase is sufficient for KHIs and AAs (Kelland et al., 2006). Generally, the best AAs can perform at higher subcooling than KHIs (Kelland, 2006). For deeper-water applications, AAs may be the only choice. A number of AAs have been developed and tested in the literature (Kelland, 2006; Kelland et al., 2006; Peng et al., 2012). A high-performance AA enables hydrates to form as a transportable flow slurry of hydrate particles dispersed in the liquid hydrocarbon phase. However, the literature on flow characteristics and rheological properties of hydrate slurry is mostly related to the aqueous phase. For example, Wang et al. (2008, 2010) investigated the flow characteristics of tetrahydrofuran (THF) and refrigerant
3

(CH3CCl2F or HCFC-141b) clathrate hydrate slurry and developed a model to determine the safe flow of hydrate slurries. Delahaye et al. (2008) studied the rheological characteristic of CO2 hydrate slurry using an experimental dynamic loop, and an empirical model was developed to describe its rheological behaviour. Delahaye et al. (2011) also studied the flow properties of CO2 hydrate slurry in the presence of additives (EO/PO copolymer). Clain et al. (2012) investigated the rheological properties of tetra-n-butylphosphonium bromide (TBPB) hydrate slurry flow for hydrate fractions between 0 and 28.2 vol% and shear rates between 100 and 700 s-1 and deduced that TBPB hydrate slurries exhibit a shear-thinning behaviour. Darbouret et al. (2005) studied the rheological properties of tetra-n-butyl ammoniumbromide (TBAB) hydrate suspensions and determined the apparent viscosity and yield shear stress for different hydrate contents. Hashimoto et al. (2011) studied TBAB and tetra-n-butylammonium fluoride (TBAF) hydrate slurries and showed that both systems present pseudoplastic behaviour. They also studied the effect of the surfactant on the flow properties of TBAB and TBAF hydrate slurries. Suzuki et al. (2013) studied the flow and hear transfer characteristics of ammonium alum hydrate slurries. Recently, Joshi et al. (2013) presented a detailed analysis of hydrate formation experiments performed in a 95-m-long flowloop (9.7-cm internal pipe diameter) in high-water-cut systems. They proposed a hydrate plugging formation mechanism, which involves the transition from a homogeneous suspension (region I) of hydrate particles to a heterogeneous suspension (region II), leading to increased particle interaction and agglomeration and ultimately causing the formation of a hydrate bed
4

and wall deposit (region III). Gas and oil generally coexist in multiphase transportation pipelines. However, there are only a few reports of studies on the flow characteristics and rheological properties of hydrate slurries formed from the liquid hydrocarbon phase. Using a high-pressure rheology apparatus, Webb et al. (2012) studied the in situ formation and flow properties of gas hydrates from a water/crude oil emulsion. Fidel-Dufour et al. (2006) investigated the crystallisation and rheology of a methane/water/dodecane system and demonstrated that it behaves as a Newtonian fluid. According to the studies on the rheological and flow properties of gas hydrate suspensions, Sinquin et al. (2004) demonstrated that hydrate particle formation in the liquid phase modifies the flow properties and that the pressure drop is controlled by the friction factor under turbulent flow conditions or the apparent viscosity under a laminar flow regime. Shi et al. (2011) and Gong et al. (2010) investigated natural gas hydrate formation and growth at different water cuts for a water-in-condensate oil emulsion in a flow loop. Zylyftari et al. (2013) studied the salt effects on the rheological properties of a hydrate-forming emulsion. Recently, in our group, Peng et al. (2012) investigated the flow characteristics, shutting-down/restarting behaviour, and morphology of hydrate slurries formed from a (natural gas + diesel oil/condensate oil + water) system containing an anti-agglomerant. Based on the analysis of the rheology parameters and apparent viscosity of hydrate slurry during the formation of gas hydrate, they declared that the hydrate slurry exhibits shear-thinning behaviour and is a pseudoplastic fluid. However, the range of shear rate studied in their work (Peng et al., 2012) was only
5

from 120 to 360 s-1, which is insufficient to accurately describe the rheology of hydrate slurry for a wider range of shear rates. In addition, the restarting effect from the static state is important for flow safety assurance in the pipeline. Peng et al. (2012) studied the restarting effect of hydrate slurry, but the shutting-down time only ranged from a few minutes to less than 1 h. In gas/oil transport pipelines, the water cut is usually less than 30 vol%. In this work, the flow characteristics and rheological properties of hydrate slurry in a flow loop were examined for initial water cuts from 5 to 30 vol% and shear rates from 50 to 350 s-1. A new type of anti-agglomerant, different from that of Peng et al. (2012), was added to these systems. The flow rate and pressure drop of the hydrate slurry formed were systematically investigated. The morphologies of the hydrate slurry at different stages were recorded and analysed. Combined with the experimental data, an empirical rheological model based on Herschel-Bulkley-type equation was proposed to describe the rheological behaviour of hydrate slurry. In addition,

shutting-down/restarting tests with three different shutting-down times (2 h, 4 h, and 8 h) were performed for all water-cut systems to investigate the rheological properties of the hydrate slurry.

2. Experimental
2.1. Materials and apparatus The experimental materials include water, diesel oil, natural gas, and an anti-agglomerant. Diesel oil, with a freezing point of 253.2 K, serves as the liquid oil phase, and its composition is shown in Table 1, as analysed by a crude oil true boiling
6

point (TBP) distillation system. The natural gas used is the associated gas from an oilfield, and its composition was analysed by a HP7890A gas chromatograph and listed in Table 2. An anti-agglomerant patented by Chen et al. (2011) was chosen in this work, and its performance was assessed using a high-pressure sapphire cell. It is well known that the produced water contains salt; therefore, a 0.81 wt% NaCl aqueous solution was prepared and used as the aqueous phase in the experiments. The experimental flow loop illustrated in Figure 1, similar to that described in our previous work (Peng et al., 2012; Shi et al., 2011), was used to measure the flow characteristics and rheological properties of natural gas hydrate slurries with six different initial water cuts from 5.0 to 30.0 vol%. It mainly consists of a U-bend double pipe (20 m long, 25.4-mm inner diameter) made of 316L stainless steel, with a maximum operation pressure of 10.0 MPa. The pipes were maintained at constant temperature with two fluid circulation baths (Neslab RTE 111D). Five thermocouples (0.1 K) and a pressure gauge (0-10 MPa, 0.1%) were adopted to measure the temperature and pressure during the experiment, respectively. An IH-type single-stage, single-suction, cantilever centrifugal pump (Tianjin Pumps & Machinery Group Co., Ltd., China) was equipped to circulate the liquid through the pipe loop. A turbine flow meter was used to measure the volumetric flow rate, and a differential pressure transducer was placed between the inlet and outlet of the U-bend pipe to measure the pressure drop generated by the fluid flow. An observation window was placed in the middle of the flow loop to observe the variation of the morphology of the fluid. A mixing tank (approximately 20 L) was used to separate the gas and liquid phases, in
7

which a funnel shape in the bottom was designed to ensure that the fluid in the mixing tank and all of the slurry would be transported to the loop. All of the sensors were connected to a PC-based acquisition system. 2.2. Experimental procedure The experimental procedure for investigating the flow characteristics and rheological behaviour of hydrate slurry at fixed temperature and pressure is described as follows. First, the flow loop was cleaned by flushing with a detergent and pure water in cycles and drained using hot gas. Thereafter, a given amount of (diesel oil + water + anti-agglomerant) fluid was charged into the mixing tank and loop pipe. The flow loop was evacuated for 90 min to remove the residual air. An initial flow rate of approximately 1.5 m3/h was applied, and the temperatures of two circulating baths were cooled to the fixed experimental value (274 K in this work). After the temperature was stable, a fixed quantity of liquefied petroleum gas was slowly injected into the mixing tank to saturate the fluid phase. After the system had stabilised, the original natural gas was injected into the mixing tank until the pressure reached the experimental value (2.10 MPa in this work). To sustain the system at a constant pressure, natural gas was continuously charged to compensate for its consumption due to hydrate formation, and this process may last for several hours. The amount of gas charged into the loop was recorded online using a mass flow meter. The variations of flow rate and pressure drop with time were also recorded. When no gas was added, the hydrate slurry could be considered to be stable in the loop. The flow rate was then changed, and the corresponding pressure drop was measured to
8

determine the rheological properties of the gas hydrate slurry. In addition, the equilibrium gas was sampled from the flow loop and analysed. During the entire experiment, the morphology of the hydrate slurry could be observed through the observation window and was continuously monitored using a video camera. Because hydrate plugs easily form in deep-sea multi-phase transportation pipelines during shutting-down and restarting periods, it is important to study the restarting effect of hydrate slurry to further understand its flow characteristics. In general, the shutting-down time used in laboratory studies have ranged from tens of seconds to several hours (Peysson et al. 2007; Lachance et al. 2012; Estanga et al. 2008; Harun et al. 2008), whereas some field tests have used shutting-down times of several days (Frostman, et al. 2001; Fu, et al. 2001). Therefore, in this work, three different shutting-down times, 2 h, 4 h, and 8 h, were adopted for the shutting-down/restarting tests of the hydrate slurry. To investigate the restarting effect of the hydrate slurry, the pump was turned off when the slurry was under steady-flow conditions. After shutting down for several hours, the pump was restarted. The flow rate and corresponding pressure drop of the hydrate slurry at the restarting state were recorded to study the restarting effect. The variation of the morphology of the hydrate slurry during this process was also observed.

3. Results and discussion


3.1. Evaluation of the new anti-agglomerant The method adopted in this work to evaluate the anti-agglomerant is the same as that used by Peng et al. (2012). A high-pressure sapphire test system devised and
9

constructed in our group was used to perform the evaluation experiments. More details about the test system can be found in our previous papers (Chen et al., 2009; Sun et al., 2003). The change in the morphology of the hydrate slurry can be observed directly through the transparent cell wall. The anti-agglomerant performance can also be evaluated by observing whether the stirrer in the fluid can move smoothly up and down after almost all water has been converted into hydrate. Figure 2 shows the morphologies of the hydrate slurries formed from the (water + diesel oil + natural gas) system, where the initial water cuts range from 5 to 30 vol%. The composition of the original gas used is listed in Table 2. The effective dosage of anti-agglomerant added is 3.0 wt% of the water mass for each run. The maximum subcooling tested by the cooling method at constant pressure is over 20 K. As shown in Figure 2, for water cuts ranging from 5 to 30 vol%, we can see that the hydrate particles are homogeneously dispersed in the diesel oil phase and do not agglomerate after almost all of the water has been converted into hydrate. Although the hydrate slurry becomes stickier with increasing initial water cut, the stirrer could still move smoothly up and down. In addition, after the hydrate slurry was allowed to rest at the maximum subcooling without stirring for 12 h, it was found that the stirrer could be successfully restarted and that the hydrate could be redispersed into the oil phase. 3.2. Flow characteristics and morphology of the hydrate slurry The flow characteristics and morphology of hydrate slurry were investigated at six different initial water cuts of 5, 10, 15, 20, 25, and 30 vol%. In each of the six
10

experimental runs, 3.0 wt% anti-agglomerant dosages were used. The system pressure and water bath temperature were kept constant at 2.1 MPa and 274.2 K, respectively. The system with an initial water cut of 5 vol% was used as an example to present the experimental results. Figure 3 shows the variations of the flow rate and pressure drop of hydrate slurry with the elapsed time, where the zero time refers to the beginning of the charge of natural gas. The variations of the flow rate and pressure drop of the hydrate slurry within the first 5 h are shown separately in Figure 4 for clarity. It can be observed that the flow rate decreases with the formation of hydrate and then becomes stable, whereas the pressure drop first increases with increasing hydrate quantity formed, then gradually decreases with some fluctuation, and finally reaches a stable value. In particular, at the initial stage after the stabilisation of the temperature, pressure, and flow rate, a sudden temperature rise occurs (see Figure 5) when hydrate appears due to the exothermic effect of hydrate crystallisation. At the same time, a sudden increase of the pressure drop and decrease of the flow rate (See Figure 4) occur because the formation of solid hydrate changes the flow characteristics of the fluid in the flow loop. In the work of Peng et al. (2012),for an anti-agglomerant comprised of a mixture of sorbitan monolaurate and polymer esters, when the initial water cut is less than 20 vol%, the flow rate is 1.0 m3/h, corresponding to a mean fluid velocity of 0.55 m/s. However, for higher-water-cut systems (20 vol%), the mean fluid velocity is less than 0.44 m/s. Based on the experimental results in this work, hydrate slurries with the anti-agglomerant adopted can flow steadily, and the mean velocity can reach more
11

than 0.55 m/s for all six groups of experiments. The temperature of the hydrate slurry was monitored and recorded by the thermocouples placed in five different positions in the flow loop. The average value of these five temperature points was considered as the temperature of the hydrate slurry. The variation of temperature with time for different water cuts is shown in Figure 5. A sudden rise can be clearly observed when gas hydrate appears in the flow loop, and then the temperature continues to increase and reaches a maximum value. The maximum temperature is higher for higher initial water cuts: 276.5, 276.8, 277.0, 277.5, 277.8, and 278.2 K for 5, 10, 15, 20, 25, and 30 vol%, respectively. This phenomenon is attributed to the formation of more hydrate and the release of more energy for higher initial water cuts. Gradually, with the decrease of the hydrate formation rate and the convective heat transfer between the cooling medium and hydrate slurry, the temperatures tend towards that of the water bath, 274.2 K, within the first 3 h of the experiments. Thereafter, the temperatures remained constant at approximately 274.2 K. The morphology of gas hydrate slurry in the flow loop was observed through an observation window, as shown in Figure 1. Four images taken at different stages for each water cut system are shown in Figure 6: the beginning stage before hydrate formation, the stable flow stage of the hydrate slurry, the shutting-down stage, and stable flow after restarting (the shutting-down/restarting tests will be discussed in Section 3.4). The images of the first two stages and the last stage were taken during the flow conditions. To obtain a better visualisation of the morphology of the hydrate
12

slurry at the shutting-down stage, the pictures of the third stage were taken after the circulating pump had been stopped for over 1 h. As shown in Figure 6, the fluid was in the form of a water-in-oil emulsion before natural gas hydrate appeared in the loop. With the charge of the original natural gas, hydrate particles were observed through the observation window, and the amount of hydrate particles formed at the beginning stage increases obviously with the increase of the initial water cut. Although the hydrate particles are heavier than the oil phase, the fluid system is homogeneously distributed in the pipeline in the form of a slurry, which could fill the entire flow loop. The variation of the morphology during hydrate formation is similar to that reported by Peng et al. (2012) for (natural gas + diesel oil + water) systems containing an anti-agglomerant comprised of sorbitan monolaurate and polymer esters in a mixing ratio of 4:1. However, in contrast to the homogeneous hydrate slurry observed in this work, even at the 30 vol% initial water cut, the heterogeneity becomes obvious with increasing initial water cut in Peng et al.s work, especially for the systems with water cuts of 20, 22, and 24 vol%. After the circulating pump was stopped, the separation of the liquid phase and solid phase occurred due to the difference in the density between gas hydrate and diesel oil. The result is that the oil phase is at the top of the flow loop, while the hydrate phase is at the bottom. This phenomenon can be clearly seen for the 5 vol% water cut in Figure 6, indicating that the original water-in-oil emulsion structure was destroyed when almost all water was converted into hydrate. This finding implies that the hydrate slurry system investigated in this work is not an emulsion but a solid
13

dispersion system. When the pump was restarted, the hydrate slurry returned to the uniform dispersion state. From the analysis of the flow characteristics and morphologies of hydrate slurry, it can be concluded that the oil and gas can be safely transported by forming stable and flowable hydrate slurries. 3.3. Rheological properties of gas hydrate slurry After the system stabilised after hydrate formation, rheological studies on the hydrate slurries, which were composed of hydrate particles dispersed in a hydrocarbon liquid phase, were performed using the flow loop and the capillary (Ostwald) viscosimeter method. Several assumptions must be made before the hydrate slurry rheology is evaluated using the capillary viscosimeter method. Hydrate slurries must be considered as pseudo-homogeneous fluids circulating in a laminar regime in a cylindrical pipe without any wall slip. The assumption of wall slip was not checked in this work because it requires the use of different pipe sizes. The assumption of a laminar regime will be discussed later. After introducing these assumptions, the flow rate, shear stress, and shear rate can be represented at the wall by the Rabinowitsch and Mooney equation (Metzner and Reed, 1955):

Q 1 = 3 3 R w

2 w d

(1)

where w is the shear stress at the wall, which is related to the pressure drop by the following expression:

w =

D P 4 L

(2)

where D is the inside diameter of pipeline, L is the pipe length, and P is the pressure drop.
14

Differentiation of the Rabinowitsch and Mooney equation yields the following expression of the shear rate at the wall:

w =
where n is the behaviour index, defined as

8u 3n + 1 D 4n

(3)

n=

d ln w 8u d ln D

(4)

Combining eqs 2 and 3, the experimental measurements of the pressure drop and flow rate can allow the rheological behaviour of the hydrate slurry to be established:

w = f ( w )

(5)

According to eq 5, the relationship between the shear stress and shear rate allows various fluid classes to be distinguished, such as Newtonian fluids ( is proportional to ) and non-Newtonian fluids ( is not proportional to ). For a given hydrate volume fraction system, the behaviour index n, consistency index k, and yield stress 0 are identified based on the general Herschel-Bulkley (HB) equation:

w =0 + k w

(6)

As noted by Anderson and Gudmundsson (2000), the apparent viscosity of the hydrate slurry can be defined as the ratio between the shear stress and the shear rate at the wall:

app =

w w

(7)

To systematically investigate the rheological behaviour of hydrate slurry, broader flow rates were examined in this work than in Peng et al. (2012), and the
15

corresponding pressure drops were recorded simultaneously after the hydrate slurry reached the stable state. Figure 7 shows an example of the rheological measurements for a 5 vol% water cut after the slurry reaches the stable state, showing the various flow rate plateaus and corresponding pressure drops. According to eqs 2 and 4, the logarithmic relationships between

D P at different hydrate volume fractions are shown in Figure 8. The data could be 4 L

8u and D

approximately regressed by a linear curve with a slope corresponding to the behaviour index, n, denoting the difference from Newtonian behaviour. Figure 9 shows the

variation of the behaviour index with the hydrate volume fraction s . It can be observed that n decreases with increasing s from 6.17 vol% to 34.88 vol%. The following correlation for n as a function of s was established:

n = 1.0000 0.4352s 3.2395s 2

(8)

From Figure 9, it can be clearly seen that the behaviour index is always less than one and decreases with increasing hydrate volume fraction, meaning that the hydrate slurry exhibits a more typical non-Newtonian behaviour with increasing hydrate volume fraction. Figure 10 represents the relationship between shear stress w , obtained from eq 2, and w n , deduced from eqs 3 and 4. The shear stress tends to zero when the value of w n decreases to zero at different hydrate volume fractions from 6.17 to 34.88 vol%. According to the HB model (eq 6), the experimental points can be modelled by a linear curve, where the consistency index k is the slope and the yield stress 0 is the ordinate at the origin. Based on the regression results, it can be found that the
16

yield stress is negligibly different from zero:

0 = 0

(9)

Figure 11 represents the evolution of the consistency index k as a function of hydrate volume fraction obtained from the experimental data. The correlation of the consistency index k as a function of the solid volume fraction obtained from the experimental data is given by the following expression:

k = exp( 4.7798 + 0.2777s + 24.3751s 2 )

(10)

As shown in Figure 11, the consistency index k grows exponentially with the hydrate volume fraction to a given extent. This index increases rapidly when the hydrate volume fraction is higher than 18.07 vol%, which means that the apparent viscosity of the hydrate slurry increases significantly under the same conditions. This phenomenon is similar to that observed by Peng et al. (2012) and Clain et al. (2012). Furthermore, the apparent viscosity of the hydrate slurry can be expressed as follows, based on eqs 6 to 10,

app = exp( 4.7798 + 0.2777s + 24.3751s ) w


2

0.4352s 3.2395s 2

(11)

The experimental apparent viscosities of the hydrate slurry at different hydrate volume fractions and shear rates determined from eqs 2, 3, 4, and 6 and the apparent viscosity predictions obtained from eq 11 were compared and are presented in Figure 12. Figure 12 clearly shows that there is a good agreement between the experimental data and model predictions for all hydrate volume fractions. In general, the apparent viscosities of the hydrate slurry decrease with increasing shear rate, meaning that the hydrate slurry fluid is a pseudoplastic fluid with a shear-thinning behaviour in this
17

work. However, for hydrate volume fractions of 6.17 vol% and 12.2 vol%, the apparent viscosities are always lower than 8.0 mPas and decrease slightly with increasing shear rate from 120 to 350 s-1, meaning that the shear-thinning behaviour is not obvious for these two hydrate volume fractions. This phenomenon can be attributed to the behaviour indexes (see Figure 9) of these two hydrate volume fractions being so close to unity that the non-Newtonian behaviour (shear-thinning) is not obvious.

3.4. Shutting-down/restarting tests


It is well known that hydrate plugs can occur easily in deep-sea multi-phase transmission pipelines in the shutting-down and restarting stages. Therefore, tests using a long shutting-down period are essential for investigating the flow characteristics of hydrate slurry. In this work, three shutting-down times, 2 h, 4 h, and 8 h, were applied for each run, which are much longer than the shutting-down times investigated by Peng et al. (2012).Table 3 lists the flow rates and pressure drops at seven different stages for different initial water cuts: stable flow before shutting down, restarting after shutting down for 2 h, stable flow after shutting down for 2 h, restarting after shutting down for 4 h, stable flow after shutting down for 4 h, restarting after shutting down for 8 h, and stable flow after shutting down for 8 h. Table 3 clearly shows that hydrate slurry in the presence of anti-agglomerant can be easily and safely restarted after spending a long time in the static state in all experiments. The restarting effect on the hydrate slurry becomes more obvious as the shutting-down time increases from 2 h to 8 h for a given hydrate volume fraction
18

system. For instance, the pressure drop and flow rate of the hydrate slurry at the restarting stage after shutting down for 2 h for a hydrate volume fraction of 12.2% are 10.51 kPa and 1.30 m3/h, respectively, while the pressure drop and flow rate increase to 11.12 kPa and 1.32 m3/h at the restarting stage after shutting down for 8 h. In addition, the restarting effect also becomes more obvious with increasing hydrate volume fraction under the same shutting-down time. Compared with the hydrate volume fraction of 12.2% mentioned above, the pressure drop and flow rate at the restarting stage after shutting down for 2 h and 8 h for a hydrate volume fraction of 34.88% are 11.56 kPa and 1.26 m3/h and 14.13 kPa and 1.37 m3/h, respectively. Therefore, the hydrate slurry studied in this work presents a typical thixotropic behaviour, which becomes more obvious with the increase of the hydrate volume fraction. This phenomenon can be attributed to the microscopic structure of the hydrate slurry. When the hydrate slurry is subjected to a long shutting-down period, a netted texture may form due to the adhesive force between hydrate particles. When the pump is restarted, the netted structure will be destroyed. However, it may take some time to build the new stable inner structure because of the adhesive force between the hydrate particles. This time can be regarded as the reason that the hydrate slurry investigated in this work exhibits thixotropic behaviour after a long shutting-down time. As discussed in Section 3.3, the results mentioned in this work are valid when two assumptions are satisfied: laminar flow and no wall slip. According to the method used by Clain et al. (2012), in which the TBPB hydrate slurry was verified to be in the
19

laminar regime by the determination of the Metzner-Reed Reynolds number ReMR (Metzner and Reed, 1955) and the Fanning friction factor f, the first assumption can be validated. The Metzner-Reed Reynolds number is determined from the behaviour index n and consistency index k:

Re MR =

D nU 2 n HS 1 + 3n n n 1 k( ) 8 4n

(12)

If the fluids are in laminar regime, the Fanning factor and the Metzner-Reed Reynolds number can be correlated using the Hagen-Poiseuille equation regardless of whether the fluid exhibits Newtonian or non-Newtonian behaviour:

f =

16 Re MR

(13)

The classical expression of the regular Fanning factor contributions as a function of the fluid velocity and pressure drop can be written as follows:

f =

DP 2 HS LU 2

(14)

Figure 13 presents the relationship between the Fanning factor and the Metzner-Reed Reynolds number for hydrate slurry with different hydrate volume fractions. The experimental data obtained from eq 14 are in good agreement with the prediction of the Hagen-Poiseuille equation, which indicates that the hydrate slurries investigated in this work are all in the laminar regime.

4. Conclusions
A dynamic loop was adopted to investigate the flow characteristics and rheological properties of gas hydrate slurry in the presence of anti-agglomerant, where
20

the initial water cuts range from 5.0 to 30.0 vol%, and the shear rates range from 50 to 350 s-1. The experimental results demonstrate that the hydrate slurry exhibits a shear-thinning behaviour and is a pseudoplastic fluid. The slurrys non-Newtonian behaviour becomes more obvious with increasing hydrate volume fraction. An empirical HB-type equation that included solid fraction dependency was used to describe the rheological behaviour of the gas hydrate slurry. The apparent viscosities of the hydrate slurry with different hydrate volume fractions were determined by the new model and were in good agreement with the experimental data. The shutting-down/restarting tests indicate that the hydrate slurry exhibits obvious thixotropic behaviour. Based on the analysis of the flow characteristics and morphologies of the hydrate slurry, the oil and gas can be safely transported by forming stable and flowable hydrate slurries.

Acknowledgements
The financial support received from National 973 Project of China (No. 2012CB215005) and National Natural Science Foundation of China (Nos. 20925623, U1162205, 51376195) are gratefully acknowledged.

References
Andersson, V., Gudmundsson, J.S. 2000. Flow properties of hydrate-in-water slurries. Ann. N.Y. Acad. Sci. 912, 322-329. Arjmandi, M., Tohidi, B., Danesh, A., Todd, A.C., 2005. Is subcooling the right driving force for testing low-dosage hydrate inhibitors? Chem. Eng. Sci. 60, 1313-1321. Chandragupthan, B., 2011. PetroMin Pipeliner 37, 50-57. Chen, L.T., Sun, C.Y., Chen, G.J., Nie, Y.Q., Sun, Z.S., Liu, Y.T., 2009. Measurements
21

of hydrate equilibrium conditions for CH4, CO2, and CH4 + C2H6 + C3H8 in Various systems by step-heating method. Chin. J. Chem. Eng. 17, 635-641. Chen, G.J., Li, W.Z., Li, Q.P., Sun, C.Y., Mu, L., Chen, J., Peng, B.Z., Yang, Y.T., Meng, H., 2011. China Patent Number 201110096579.2. Clain, P., Delahaye, A., Fournaison, L., Mayoufi, N., Dalmazzone, D., Frst, W., 2012. Rheological properties of tetra-n-butylphosphonium bromide hydrate slurry flow. Chem. Eng. J. 193194, 112-122. Creek, J.L., 2012. Efficient hydrate plug prevention. Energy Fuels 26, 4112-4116. Darbouret, M., Cournil, M., Herri, J.M., 2005. Rheological study of TBAB hydrate slurries as secondary two-phase refrigerants. Int. J. Refrig. 28, 663-671. Delahaye, A., Fournaison, L., Marinhas, S., Martnez, M.C., 2008. Rheological study of CO2 hydrate slurry in a dynamic loop applied to secondary refrigeration. Chem. Eng. Sci. 63, 3551-3559. Delahaye, A., Fournaison, L., Jerbi, S., Mayoufi, N., 2011. Rheological properties of CO2 hydrate slurry flow in the presence of additives. Ind. Eng. Chem. Res. 50, 8344-8353. Estanga, D. A., Creek, J., Subramanian, S., Kini, R. A., 2008. Last 20 years of gas hydrates in the oil industry: challenges and achievements in predicting pipeline blockage. Proceedings of the 6th International Conference on Gas Hydrates (ICGH 2008), Vancouver, British Columbia, CANADA, July 6-10. Fidel-Dufour, A., Gruy, F., Herri, J.M., 2006. Rheology of methane hydrate slurries during their crystallization in a water in dodecane emulsion under flowing. Chem. Eng. Sci. 61, 505-515. Frostman, L. M., Przybylinski, J. L., 2001. Successful applications of

anti-agglomerant hydrate inhibitors. SPE International Symposium on Oilfield Chemistry, February 13-16, Houston, Texas. Fu, B., Neff, S., Mathur, A., Bakeev, K., 2001. Novel low dosage hydrate inhibitors for deepwater operations. SPE Annual Technical Conference and Exhibition, September 30 - October 3, New Orleans, Louisiana. Gao, S.Q., 2009. Hydrate risk management at high watercuts with anti-agglomerant
22

hydrate inhibitors. Energy Fuels 23, 2118-2121. Gong, J., Shi, B., Zhao, J., 2010. Natural gas hydrate shell model in gas-slurry pipeline flow. J. Nat. Gas Chem. 19, 261-266. Harun, A.F., Fung, G., Erdogmus, M., 2008. Experience in AA-LDHI usage for a deepwater gulf of Mexico dry-tree oil well: pushing the technology limit. SPE Prod. Oper. 23 (1), 100-107. Hashimoto, S., Kawamura. K., Ito H., Nobeoka M., Ohgaki K., Inoue Y., 2011. Rheological study on tetra-n-butyl ammonium salt semi-clathrate hydrate slurries. Proceedings of 7th International Conference on Gas Hydrates (ICGH 2011), Edinburgh, Scotland, U. K., July 17-21. Huo, Z., Freer, E., Lamar, M., Sannigrahi, B., Knauss, D.M., Sloan, E.D., 2001. Hydrate plug prevention by anti-agglomeration. Chem. Eng. Sci. 56, 4979-4991. Joshi, S.V., Grasso, G.A., Lafond, P.G., Rao, I.,Webb, E., Zerpa, L.E., Sloan, E.D., Koh, C.A., Sum, A.K., 2013. Experimental flowloop investigations of gas hydrate formation in high water cut systems. Chem. Eng. Sci. 97, 198-209. Kelland, M.A., 2006. History of the development of low dosage hydrate inhibitors. Energy Fuels 20, 825-847. Kelland, M.A., Svartaas, T.M., vsthus, J., Tomita, T., Chosa, J.I., 2006. Studies on some zwitterionic surfactant gas hydrate anti-agglomerants. Chem. Eng. Sci. 61, 4048-4059. Lachance, J. W., Talley, L. D., Shatto, D. P., Turner, D. J., Eaton, M. W., 2012. Formation of hydrate slurries in a once-through operation. Energy Fuels 26, 4059-4066. Mehta, A.P., Hebert, P.B., Cadena, E.R., Weatherman, J.P., 2003. SPE Prod. Facil. Feb, 73-79. Metzner, A.B., Reed, J., 1955. Flow of non-Newtonian fluids-correlation of the laminar, transition, and turbulent-flow regions. AIChE J. 1, 434-440. Peng, B.Z., Chen, J., Sun, C.Y., Dandekar, A., Guo, S.H., Liu, B., Mu, L., Yang, L.Y., Li, W.Z., Chen, G.J., 2012. Flow characteristics and morphology of hydrate slurry formed from (natural gas+diesel oil/condensate oil+water) system
23

containing anti-agglomerant. Chem. Eng. Sci. 84, 333-344. Peysson, Y., Bensakhria, A., Antonini, G., Argillier, J. F., 2007. Pipeline Lubrication of Heavy Oil: Experimental Investigation of Flow and Restart Problems. SPE Prod. Oper. 22 (1), 135-140. Shi, B.H., Gong, J., Sun, C.Y., Zhao, J.K., Ding, Y., Chen, G.J., 2011. An inward and outward natural gas hydrates growth shell model considering intrinsic kinetics, mass and heat transfer. Chem. Eng. J. 171, 1308-1316. Sinquin A., Palermo, T., Peysson, Y., 2004. Rheological and flow properties of gas hydrate suspensions. Oil Gas Sci. Technol. Rev. IFP 59, 41-57. Sloan, E. D. 2003. Fundamental principles and applications of natural gas hydrates. Nature 426, 353-363. Sloan, E.D., Koh, C.A., 2008. Clathrate hydrates of natural gases, 3rd ed. CRC Press (Taylor and Francis Group): Boca Raton, FL. Sloan, E.D., Koh, C., Sum, A., Ballard, A.L., Creek, J., Eaton, M., Lachance, J., McMullen, N., Palermo, T., Shoup, G., Talley, L., 2011. Natural gas hydrates in flow assurance. Gulf Professional Publishing: Oxford, U.K. Sun, C.Y., Chen, G.J., Lin, W., Guo, T.M., 2003. Hydrate formation conditions of sour natural gases. J. Chem. Eng. Data 48, 600-602. Suzuki, H., Konaka, T., Komoda, Y., Ishigami, T., 2013. Flow and heat transfer characteristics of ammonium alum hydrate slurries. Int. J. Refrig. 36, 81-87. Wang, W.C., Fan, S.S., Liang, D.Q., Yang, X.Y., 2008. Experimental study on flow characters of CH3CCl2F hydrate slurry. Int. J. Refrig. 31, 371-378. Wang, W.C., Fan, S.S., Liang, D.Q., Li, Y.X., 2010. A model for estimating flow assurance of hydrate slurry in pipelines. J. Nat. Gas. Chem. 19, 380-384. Webb, E.B., Rensing, P.J., Koh, C.A., Sloan, E.D., Sum, A.K., Liberatore, M.W., 2012. High-pressure rheology of hydrate slurries formed from water-in-oil emulsions. Energy Fuels 26, 3504-3509.

Zylyftari, G., Lee, J.W., Morris, J.F., 2013. Salt effects on thermodynamic and rheological properties of hydrate forming emulsions. Chem. Eng. Sci. 95, 148-160.
24

Table 1. Composition of the diesel oil used


component heptanes octanes nonanes decanes undecanes dodecanes tridecanes tetradecanes pentadecanes hexadecanes heptadecanes octadecanes eicosanes tetracosanes octacosanes plus total mol% 0.50 0.50 2.81 7.74 8.74 9.95 8.74 6.53 4.92 4.72 5.33 6.83 14.47 15.78 2.41 100.00 wt% 1.05 0.92 4.60 11.40 11.73 12.24 9.94 6.90 4.86 4.37 4.64 5.63 10.74 9.77 1.28 100.00

25

Table 2. Composition of the original natural gas used


component methane ethane propane i-butane n-butane i-pentane n-pentane carbon dioxide nitrogen Total mol% 85.41 6.01 5.79 0.16 0.05 0.01 0.02 0.02 2.53 100.00

26

Table 3. Variation of the flow rate and pressure drop during shutting-down/restarting tests for different initial water cuts
stage stable flow before shutting down restart after 2h shutting down stable flow after restarting restart after 4h shutting down stable flow after restarting restart after 8h shutting down stable flow after restarting

hydrate volume fraction ( vol%) 6.17 pressure drop (kPa) flow rate (m /h) pressure drop (kPa) flow rate (m /h) pressure drop (kPa) flow rate (m /h) pressure drop (kPa) flow rate (m /h) pressure drop (kPa) flow rate (m /h) pressure drop (kPa) flow rate (m /h) pressure drop (kPa) flow rate (m /h)
3 3 3 3 3 3 3

12.2 8.32 1.23 10.51 1.30 8.35 1.23 10.75 1.25 8.34 1.21 11.12 1.32 8.34 1.22

18.07 8.56 1.23 10.58 1.26 8.50 1.20 11.21 1.26 8.49 1.22 12.63 1.36 8.72 1.21

23.81 9.25 1.21 10.91 1.30 9.17 1.19 11.33 1.34 9.32 1.21 13.49 1.42 9.28 1.23

29.4 9.91 1.19 10.97 1.25 9.55 1.16 11.95 1.31 9.47 1.15 13.75 1.35 9.73 1.16

34.88 9.98 1.15 11.56 1.26 9.31 1.14 12.15 1.32 9.27 1.15 14.13 1.32 9.93 1.13

7.97 1.25 10.20 1.29 7.79 1.28 10.11 1.26 7.91 1.25 10.49 1.42 7.98 1.27

27

Figure Captions
Figure 1. Schematic of hydrate flow-loop system. Figure 2. Morphologies of natural gas hydrate slurry formed in the high pressure sapphire cell with six different initial water cuts when at 274.2 K and 7.50 MPa. Figure 3. Variation of the fluid flow rate and pressure drop of the hydrate slurry with the elapsed time at 5 vol% water cut. Figure 4. Variation of the fluid flow rate and pressure drop of the hydrate slurry within the first 5 h at 5 vol% water cut. Figure 5. Variation of temperature of the hydrate slurry with time at different initial water cut. Figure 6. Morphologies of the emulsion or the hydrate slurry at different water cuts and stages. Figure 7. Variation of pressure drop by adjusting the flow rate during the measurement of the rheological behaviour of the hydrate slurry when at 5 vol% water cut. Figure 8. Logarithmic relationship between 8uav D and D P 4 L for the hydrate

slurry at different hydrate volume fractions. Figure 9. Behaviour index as a function of the hydrate volume fraction. Figure 10. Shear stress w as a function of shear rate w fractions from 6.17 to 34.88 vol%. Figure 11. Variation of consistency index with the hydrate volume fraction. Figure 12. Comparison of apparent viscosity of the hydrate slurry between experimental data and model prediction for different hydrate volume fractions. Figure 13. Variation of friction factor of the hydrate slurry as a function of Reynolds number for different hydrate volume fractions.

n

for hydrate volume

28


Resistance thermocouple detector Scanning Thermometer Recirculation Tube Differential Pressure Transducer

Hydrate Tube Swagelok Union

Flowmeter

P Visual Window mixing tank

Pump

Data Acquistion System Cooling System Gas Cylinder Circulating Bath U-bend

Figure 1. Schematic of hydrate flow-loop system.

29

5 vol% water cut 10 vol% water cut

15 vol% water cut

20 vol% water cut 25 vol% water cut

30 vol% water cut

Figure 2. Morphologies of natural gas hydrate slurry formed in the high pressure sapphire cell with six different initial water cuts when at 274.2 K and 7.50 MPa.

30

1.8
hydrate formation

12 flow rate Pressure drop 10 8 Pressure drop (kPa)

1.5 Flow rate (m /h) 1.2 0.9


3

6 0.6 0.3 0.0 4 2 25

10 Time (h)

15

20

Figure 3. Variation of the fluid flow rate and pressure drop of the hydrate slurry with the elapsed time at 5 vol% water cut.

31

1.8
hydrate formation

1.5 Flow rate (m /h) 1.2 0.9


3

flow rate Pressure drop


9

0.6 0.3
3

0.0

2 Time (h)

Figure 4. Variation of the fluid flow rate and pressure drop of the hydrate slurry within the first 5 h at 5 vol% water cut.

32

Pressure drop (kPa)

278 Temperature (K) 277 276 275 274 0


hydrate formation

initial water cut


5 vol% 10 vol% 15 vol% 20 vol% 25 vol% 30 vol%

2 Time (h)

Figure 5. Variation of temperature of the hydrate slurry with the time at different initial water cut.

33

5 vol% water cut

10 vol% water cut

15 vol% water cut

20 vol% water cut

25 vol% water cut

30 vol% water cut Before the formation Stable flow Shutting down After restarting

Figure 6. Morphologies of the emulsion or the hydrate slurry at different water cuts and stages.

34

12

1.5

flow rate Pressure drop


9

1.0
6

0.5
3

19.6

20.0

20.4 Time (h)

20.8

21.2

Figure 7. Variation of pressure drop by adjusting the flow rate during the measurement of the rheological behaviour of the hydrate slurry when at 5 vol% water cut.

35

Pressure drop (kPa)

Flow rate (m /h)

1.6 1.4 1.2 ln(DP/4/L) 1.0 0.8 0.6 0.4 0.2 0.0

Hydrate volum fraction


6.17 vol% 12.20 vol% 18.07 vol% 23.81 vol% 29.41 vol% 34.88 vol% Regressed

3.3

3.6

3.9

4.2 4.5 4.8 ln(8uav/D)

5.1

5.4

5.7

Figure 8. Logarithmic relationship between 8uav D and D P 4 L for the hydrate slurry at different hydrate volume fractions.

36

1.0 0.9 Behaviour index 0.8 0.7 0.6 0.5 0.4

Experimental data Regressed line

10 15 20 25 30 Hydrate volume fraction (vol%)

35

40

Figure 9. Behaviour index as a function of the hydrate volume fraction.


37

6 5 4
w (Pa)

Hydrate volume fraction


6.17 vol% 12.20 vol% 18.07 vol% 23.81 vol% 29.41 vol% 34.88 vol% Regressed line

3 2 1 0 0 50 100 150

(s-n)

200

250

300
n

350

Figure 10. Shear stress w as a function of shear rate w fraction from 6.17 to 34.88 vol%.

for hydrate volume

38

0.21 Consistency index k (Pa.s )


n

0.18 0.15 0.12 0.09 0.06 0.03 0.00 5 10

Experimental Data Regressed line

15 20 25 30 Hydrate volume fraction (vol%)

35

40

Figure 11. Variation of consistency index with the hydrate volume fraction.

39

22 20 18
app (mPas)

Hydrate volum fraction


6.17 vol% 12.20 vol% 18.07 vol% 23.81 vol% 29.41 vol% 34.88 vol% Model line

16 14 12 10 8 6 40 80 120 160 200

w (s-1)

240

280

320

360

Figure 12. Comparison of apparent viscosity of the hydrate slurry between experimental data and model prediction for different hydrate volume fractions.

40

100

10 Fanning factor

1
Hydrate volume fraction
6.17 vol% 12.20 vol% 18.07 vol% 23.81 vol% 29.40 vol% 34.88 vol% 16/ReMR

0.1

0.01

1E-3

10

ReMR

100

1000

Figure 13. Variation of friction factor of the hydrate slurry as a function of Reynolds number for different hydrate volume fractions.

41

Higlights

Hydrate slurry presents obvious shearthinning behaviour with increase of hydrateratio.

HydrateslurryistransportedasasoliddispersionsystemwithadditionofAAs. A HerschelBulkley type equation was built by considering the hydrate volume fraction.

Shuttingdown/restarting tests show that the hydrate slurry is easily and safely restarted.

Hydrate slurry exhibits obvious thixotropic behaviour with increasing shuttingdowntime.

42

You might also like