Next Article in Journal
NIR Spectroscopy as an Alternative to Thermogravimetric Analyzer for Biomass Proximate Analysis: Comparison of Chip and Ground Biomass Models
Previous Article in Journal
Test Methods for Validating Inductive Transformer Performance during Cable Discharge: Main Principles, Parameters and Testing Procedures
Previous Article in Special Issue
A Composite Framework Model for Transient Pressure Dynamics in Tight Gas Reservoirs Incorporating Stress Sensitivity
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Comprehensive Review of Nanotechnology Applications in Oil and Gas Well Drilling Operations

by
Alhaitham M. Alkalbani
1,2,* and
Girma T. Chala
1,*
1
Department of Mechanical Engineering (Well Engineering), International College of Engineering and Management, P.O. Box 2511, CPO Seeb, Muscat 111, Oman
2
Department of Mechanical Engineering, James Watt School of Engineering, University of Glasgow, Glasgow G12 8QQ, UK
*
Authors to whom correspondence should be addressed.
Energies 2024, 17(4), 798; https://doi.org/10.3390/en17040798
Submission received: 17 December 2023 / Revised: 9 January 2024 / Accepted: 5 February 2024 / Published: 7 February 2024
(This article belongs to the Special Issue Advances in Oil and Gas Well Engineering Science and Technology)

Abstract

:
The field of nanotechnology has shown promise in addressing major problems and improving drilling effectiveness. An overview of the difficulties encountered during oil and gas well drilling operations and the demand for creative solutions opens the debate. This review explores how nanotechnology is transforming the oil industry and enhancing performance as a whole. The evaluation of the uses of nanotechnology for better oil recovery, real-time monitoring, innovative materials, drilling fluids, and reservoir characterization are extensively discussed in this review. The primary function of additives is to improve the fundamental characteristics of drilling fluids. The variety of fluid additives available is a reflection of the complex drilling–fluid systems that are currently being used to enable drilling in increasingly difficult subsurface conditions. Common additives used in water- and oil-based drilling fluids include lubrication, shale stability, filtration control, rheology control, viscosification, and pH regulation. Drilling fluids frequently contain filtration control additives such as starch, polyanionic cellulose (PAC), carboxymethyl cellulose (CMC), and nanoparticles (NP). Commonly used rheology-modifier additives are xanthan gum, carboxymethyl cellulose, guar gum powder, and, more recently, salt-responsive zwitterionic polymers that were used as viscosifiers to water-based drilling fluids. The three main additives that regulate pH are citric acid monohydrate, potassium hydroxide, and sodium hydroxide. Additives that stabilize shale, such as potassium and sodium salts and asphaltenes, are often used. A wide range of materials are included in the category of lubricating additives, including polymers, asphaltenes, glass beads, oils of various grades, and oil-surfactants. Various fibrous materials, including wood, cotton, fibrous minerals, shredded tires from vehicles, and paper pulp, are used as additives to control circulation. Furthermore, shredded cellophane, bits of plastic laminate, plate-like minerals like mica flakes, granulated inert materials such as nut shells, and nano-polymers are used in wellbores to reduce fluid loss. The incorporation of nanoparticles into drilling fluids has produced upgraded fluids with better features, including improved lubricity, thermal stability, and filtering capacities. These developments aid in lowering friction, enhancing wellbore stability, and enhancing drilling efficiency. This paper also emphasizes how nanotechnology has made enhanced drilling equipment and materials possible. Drilling equipment’s longevity and performance are increased by nanocomposite materials that have been reinforced with nanoparticles due to their improved mechanical strength, wear resistance, and thermal stability. Advanced reservoir characterisation tools, including nanoparticle tracers and nanoscale imaging methods, can help locate the best drilling sites and increase production effectiveness. On the other hand, nanofluids and nanoemulsions can potentially increase oil recovery because they enhance fluid mobility, lower interfacial tension, and alter rock wettability. Although nanotechnology has many advantages, there are also issues that need to be resolved. For an implementation to be effective, factors including nanoparticle stability, dispersion, and potential environmental effects must be carefully taken into account. This review highlights the need for future research to create scalable manufacturing procedures, improve nanoparticle behaviour, and determine nanomaterials’ long-term environmental effects. In conclusion, this in-depth analysis illustrates the use of nanotechnology in transforming the process of drilling oil and gas wells.

1. Introduction

Three systems work simultaneously when using rotary drilling to bore a hole: the rotating system to rotate the drill bit; the lifting system, which raises and lowers the drill string into the hole; and the circulating system to transport the drilling fluid from the drill stem to the drill bit and back up to the surface hole [1]. Drilling fluids plays a critical role in guaranteeing drilling success by improving oil recovery and cutting down on the time needed to achieve early oil production [2]. Drilling fluids are comparable to blood in the physical makeup of the human body throughout the drilling operation. The drilling fluid’s transportation of the cuttings from the borehole is analogous to the blood’s removal of foreign objects from the body, and the mud pump functions as the system’s heart [3]. Moreover, in this instance, the mud-cleaning mechanism serves as the kidney and the lungs. The potential of nano-fluids in heat transmission, gel formation, drag reduction, binding for sand consolidation, wettability modification, and corrosive control has been brought to light by recent investigations.
The process of inserting nano-sized particles into a fluid at tiny volumetric fractions improves the fluid’s mechanical, optical, thermal, and rheological characteristics, resulting in nano-fluids [4]. Fluids containing nanoparticles also have the advantage of having tailored mechanical, electrical, optical, thermal, rheological, mechanical, and magnetic properties based on size and shape, in addition to improved stability against sedimentation [5]. These properties can be engineered during the manufacturing process and frequently outperform those of the base material. Restrictions still exist even when additives are added to enhance the performance of standard drilling fluids [6]. A major disadvantage is the propensity of water-based drilling fluids (WBDFs) to dissolve salts and perhaps result in unintentional density fluctuations. Furthermore, WBDFs can obstruct the passage of gas and oil through porous materials. The propensity of water-based drilling fluids (WBDFs) to cause clays to break up and scatter, as well as their inability to penetrate water-sensitive shale, are further limitations. Additionally, WBDFs may cause iron parts like drill bits, collars, and pipes to rust [7]. Oil-based drilling fluids also have drawbacks, most notably their high cost across the board. This form of mud has expensive component parts, and treating and discarding cuttings raises the overall cost. On the other hand, this class of drilling fluids presents environmental hazards because their disposal may contaminate land, contaminate aquifers that supply water, and damage coral reefs. They are also inappropriate for use in dry gas reservoirs [8].
Drilling efficiency is negatively impacted by environmental degradation, with over 1 million tons of drilling fluid waste produced annually in China’s oil and gas sectors alone [9]. Current drilling fluids are inappropriate for direct disposal due to strict environmental protection rules and norms, emphasizing the need for environmentally responsible drilling fluids that naturally deteriorate or disintegrate when released into the environment [10]. Water-based drilling fluids (WBDFs) are considered more environmentally friendly than oil-based drilling fluids (OBDFs). Still, they are less effective at preventing common formation damage problems during unconventional oil and gas drilling [3,11]. Nanotechnology offers potential solutions to improve drilling efficiency, boost production rates, and reduce environmental effects in the oil and gas industry [12]. However, the current state of nanotechnology applications in drilling operations is not well documented. This review aims to address this gap by examining the potential of nanotechnology in the drilling process for oil and gas wells. The analysis explores current applications, challenges, advantages, and the future potential of nanotechnology in the field. It also aims to improve the breadth and depth of the existing literature by filling in these gaps. The main topic of discussion covers the difficulties and solutions that come with drilling gas and oil wells, emphasising the revolutionary effects of nanotechnology in the oil field. Relevant statistics about the use of nanoparticles are also presented to provide a quantitative viewpoint.

2. Nanotechnology Revolutionizing the Oil Field

An artificial or naturally occurring nanoparticle with a size between 1 and 100 nm may be used in nanotechnology for a variety of applications [13]. Due to their special qualities, nanoparticles are useful in various applications in the oilfield, such as sensing or imaging, improved oil recovery (EOR), gas mobility control, drilling and completion, produced fluid treatment, and tight reservoir applications [14,15]. The articles published over the past 20 years on the use of nanoparticles (NPs) in drilling are shown in Figure 1. Interestingly, a significant rise in publications has been noted, and this trend has continued into the current research year. From 2003 to 2009, the number of research articles published annually was fewer than 10, suggesting that the use of nanotechnology in drilling fluids was not given much attention during that time. On the other hand, between 2010 and 2012, the output of publications doubled and tripled, respectively. With 79 publications, the highest number of articles on nano-drilling fluids was seen in 2020.
Nanoparticles can enter formation pores closed off to larger particles, and their size-dependent material properties can be tailored for specific applications, enhancing their efficiency. The surface-to-area ratio of nanoparticles is remarkably high, further increasing their efficiency. Current research focuses on how nanoparticles work in reservoirs, but nanoparticle propagation, delivery, and recovery have received scant attention. Additionally, the lack of apparent health, safety, and environmental guidelines for the safe distribution and recovery of nanoparticles presents a serious problem that necessitates additional focused research [17]. The affordability of nanoparticles is another challenge due to the vast amounts required for oilfield applications and the limited number of vendors [18]. Research and development focusing on the use of naturally occurring and industrial waste nanoparticles for oilfield purposes is suggested to overcome cost challenges. Imaging, drilling in unstable zones, tight reservoir applications, and EOR are identified as having the largest potential impact among the six application categories, revolutionizing planning for field development and boosting stability, efficacy, cost-effectiveness, and environmental sustainability.
Nanotechnology has been successfully applied in various industries; however, its use in the oil and gas sector, particularly in upstream exploration and production, is still limited [19]. Nanotechnology is viewed as a viable means of overcoming technological obstacles in the sector, and governments and the global oil industry have increased funding for nanotechnology research for precise reservoir measurements, nanofluids to increase oil production, nanocatalysts for oil refining, and nanomembranes for effective oil, water, and gas separation [20]. In secondary and tertiary oil recovery operations, nanotechnology has the potential to significantly improve oil recovery and address problems caused by formation damage during water and gas reservoir flooding. Nanotechnologies can improve fluid phase separation, subsurface porous media qualities, coatings for reservoir components, and enhance the functionality of manufacturing system sensors and controls [21]. Janus nanoparticles (JNPs) are emphasized for their stability compared to other types of nanoparticles in EOR applications, offering promise for carbon capture and storage (CCS) through enhanced CO2 reservoir flooding [22]. NP-enhanced CO2 reservoir flooding, using nanoparticles in liquid formulations with surfactants or polymers, CO2 nanofoams, and nanoemulsions, can enhance hydraulic fracturing, alter reservoir wettability, lower interfacial tension, guard against formation damage, and stop asphaltene precipitation [22].
Essentially, nanoparticles have unique characteristics, primarily their size and far higher surface-to-area ratio than their larger counterparts. This property increases their reactivity and interaction with surrounding surfaces, improving the conveying fluid’s qualities with a smaller amount of material [23]. Furthermore, because of their tiny size, they may easily fit through tiny pores in formations, allowing for smooth circulation inside these areas. Therefore, these remarkable characteristics provide justification for the use of nanoparticles in the oil and gas industry. The distribution of the various nanoparticles examined in the reviewed literature is depicted in Table 1. It is evident that silicon dioxide and aluminium oxide are the two most commonly utilized nanoparticles in the oil and gas sector.
Table 2 illustrates the most coveted property for nanoparticle enhancement. The most extensively researched attribute happens to be the lubricity of drilling fluids. Nanoparticles have undergone comprehensive examination to assess their impact on augmenting oil recovery, particularly within EOR applications. This is closely followed by endeavours to enhance the filtration traits of drilling fluids and improve cementing processes [25].
The increased properties of nanomaterials that incorporate nanoparticles into their structures are inherited from the nanoparticles. As nanoparticles are included in these materials’ structure, strengthening is provided. Atoms are often found inside traditional materials, while a significant number of atoms are found on the surface of nanomaterials. Nanomaterials can display exceptional chemical, optical, mechanical, electrical, thermal, and magnetic capabilities thanks to surface enrichment. Numerous techniques, such as chemical vapor deposition, plasma arcing, electrodeposition, sol–gel synthesis, ball milling, and utilising naturally occurring nanoparticles, can be used to create nanomaterials [26]. Chemical vapour deposition is the process of creating nano-particulate materials from the gas phase, where the substance is first vaporized and then, under carefully controlled circumstances, deposited as a solid on a surface. Carbon nanotubes can be created through plasma arcing, which requires ionizing gas to create plasma. In order to precisely manage the deposition of nanoparticles, electrodeposition enables the controlled deployment of one or more layers on a surface. The process of creating colloidal suspensions (sol) and then allowing them to gel to create a network structure is known as sol–gel synthesis. Small balls are used in ball milling to strike and rotate a solid substance, dissolving it into nanocrystallites. This approach successfully disperses oxide particles uniformly while generating fine particles. Last but not least, reforming already-existing nanoparticles into fresh materials entails severing their crystallite linkages. Nanomaterials are used in many different types of products, including insulation, machine tools, batteries, high-power magnets, automobiles, airplanes, and medical implants [27].
Scientists have entered the enormous field of study known as nanotechnology to explore and improve material qualities at the atomic level. Materials display distinctive properties in the nanoscale, which covers dimensions between 1 and 100 nm, and these qualities provide important insights for enhancing current materials and creating brand-new ones with exceptional features. Successful applications of nanotechnology in many facets of life have boosted the effectiveness of product development and use. In scientific contexts, the term “nano” denotes a billionth or a factor of 10−9. The unit of spatial measurement known as a nanometre is therefore one billionth of a meter [28].
Ko and Huh, 2019 [18], discussed that nanotechnology is the study of nanoscale phenomena, the practice of nanoscale engineering, and the use of materials at the nanoscale. Dimensions of the order of 100 nm or less are referred to as being on the nanoscale. The subject of materials science has seen a spectacular transformation thanks to nanoscience, which has improved analytical instruments that function at the nanoscale and opened up a wide range of applications. In a 1959 speech to the American Physical Society, physicist Richard Feynman initially presented the idea of nanoscience and nanotechnology by discussing the potential for manipulating and directing materials at the nanoscale [29].
Numerous aspects of society have been impacted by nanotechnology, leading to increased productivity and lower cost production of high-quality goods. In turn, this has caused the demand to rise. Consequently, nanotechnology can be viewed as an essential component of contemporary existence. Incorporating nanotechnology, also known as nano-tech, has aided in creating improved and effective goods in all spheres of society. Nanotechnology is currently used in various industries, such as electronics, materials and manufacturing, aircraft, photo-topography, construction, chemicals, and petroleum refineries. Chemical processes have been facilitated in the oil and gas industries, from upstream to downstream, using materials with special size-dependent characteristics [30].
Nanomaterials are the building blocks of nanotechnology, and the impacts of the nanoscale are what give them their unique properties [31]. Therefore, accurate nanomaterial characterization is essential for nanoscale research. Advances in nano-characterization technology constantly drive the advancement of new nanotechnology. Reservoirs contain a large number of micro-nano mineral particles, pores, and organic clusters because they are made of porous materials of different mineral particles arranged in a specific configuration with high heterogeneity. Reservoirs can be viewed on a microscopic scale as intricate natural nanomaterial formations. The main topics of reservoir research are the examination of reservoir space, the presence of residual fluids, and the distribution of organic materials. In order to study and test the material composition, structure, and characteristics of things at the nanoscale, nano-characterization technology is used, which stimulates research into such materials and the creation of testing instruments and procedures [32].
The amount of research carried out on nanoparticles for the four previously described applications in the oil and gas industry is shown in Figure 2. It appears that researchers have concentrated on investigating the use of nanoparticles in cementing, drilling, and EOR applications. Nonetheless, the application of nanoparticles for stimulation purposes is still rather limited compared to other uses [33].

3. Challenges Faced in Oil and Gas Well Drilling Operations

Throughout the drilling process, the accurate knowledge of the characteristics of the drilling fluid under downhole circumstances is essential [6]. The rheology of drilling fluids is considerably impacted by the elevated temperatures [34]. For instance, clay swelling, flocculation, and sodium ion replacement in bentonite mud, frequently used in drilling, cause increased yield stress at higher temperatures [35]. Segeev et al., 2019 [36], argued that high temperatures could also destroy polymeric additives in drilling fluids, diminishing viscosity and performance and creating new difficulties for drilling operations. Although high pressure alters the volume of the continuous phase of oil-based drilling fluids through compression, its effect is less pronounced than temperature, particularly under high-temperature conditions [37]. Changes in the composition of drilling mud brought on by the contamination of drilled cuttings and formation fluids can impact the rheology of drilling mud during drilling operations [38,39]. Bageri et al., 2020 [40], used rock samples from various sandstone formations, such as calcareous, arenite, ferruginous, and argillaceous formations, to perform a study to examine the effect of drilled cuttings on mud characteristics. Drilled cuttings were added at weights of 15% and 30% of the total solid weight in the mud system. The investigation concluded that various drilled cuttings made the drilling fluid more viscous. It was discovered that after 10 min, the apparent viscosity, plastic viscosity, yield point, and gel strength increased by 50–139%, 20–113%, 50–161%, and 1–26%, respectively. Due to their high clay content (30% by weight), argillaceous formations significantly affected drilling fluid characteristics more than other formations. Therefore, increasing the clay concentration in the drilled formation causes the drilling fluid’s rheological characteristics to be higher.
Drilling fluid is used during the wellbore drilling process to cause a volumetric displacement of rock [41,42]. The drilling fluid’s required density depends on the pressure within the subterranean formation. A redistribution of stress occurs along the wellbore wall due to the natural difference in density between the rock and the drilling fluid. Subsurface rock formations undergo an in situ compressive stress condition prior to drilling operations, which is typified by vertical stress (σv), minimum horizontal stress (σh), and maximum horizontal stress (σH). Following the drilling operation, a redistribution of stress occurs, resulting in modifications to the tangential, axial, and radial stress components. It is noteworthy that, notwithstanding stress modifications on the wellbore wall, rock situated at a distance from the wellbore retains its initial in situ stresses, as illustrated in Figure 3. The density of the drilling fluid and the consequent pressure exerted are critical parameters, as deviations from optimal values may instigate hydraulic fracturing or breakout, serving as dominant mechanisms for rock failure during the drilling process. In directional or horizontal drilling, the stability of the wellbore is intricately linked to its orientation in relation to the initial in situ stress conditions [43].
The investigation by Mukhametshin et al., 2017 [45], on the effect of formation lithology on drilling mud filtration and plastering was part of a larger study. They evaluated various core samples utilizing dynamic-radial filtration to mimic actual field circumstances. The lithology, temperature, differential pressure, pipe eccentricity, fractures, and lost circulation materials (LCMs) were some variables considered during the filtration trials. Lithology is highlighted to be essential in the effectiveness of dynamic filtration. It was discussed that filtration performance testing in labs employing homogeneous ceramic discs, which do not closely mimic filtration behaviour in natural formations, can produce false results. Drilling mud must be properly designed; its properties need to be continuously monitored; effective mitigation techniques must be put in place; and prompt action should be taken to reduce drilling fluid rheology-related problems in drilling operations. Shale is a type of sedimentary rock that is mostly made up of reactive and non-reactive clay minerals, including smectite, illite, kaolinite, chlorite, and vermiculite. Reactive clay minerals in shale formations encounter a number of difficulties when exposed to water, particularly wellbore instability, which can impact drilling and fracturing procedures. According to Nagarajan et al., 2021 [46], clay minerals are made up of flaky crystalline platelets with several unit layers of tetrahedral or octahedral sheets joined by mineral fragments like silicon, oxygen, and aluminium atoms. The unusual interaction of silicon and aluminium atoms with oxygen results in the formation of silica tetrahedron (-T-) sheets and alumina octahedron (-O-) sheets. Four oxygen atoms surround metallic cations to form a hexagonal, continuous two-dimensional layer known as a tetrahedral sheet. Fe3+, Al3+, or Si4+ are common metals found in tetrahedral sheets. In contrast, an octahedral sheet has a metallic cation bound to six oxygen atoms that are connected to the octahedron via shared edges in the middle. Fe2+, Fe3+, Mg2+, or Al3+ are common cations found in octahedral sheets [47].
The necessity for self-sufficiency in oil and gas resources has arisen as a result of China’s growing reliance on foreign oil, which has reached a level that raises international security concerns [48]. Shale, tight oil/gas, and coalbed methane reserves account for more than 75% of China’s total national oil and gas reserves, making them essential alternative resources [49]. However, the use of drilling fluids in unconventional oil and gas reservoirs can result in serious problems that impact productivity and efficiency [50]. It has been demonstrated that drilling fluids can change the physical and chemical characteristics of reservoirs, obstruct oil and gas flow channels, lower production rates, result in the loss of new oil and gas fields, and even harm oil and gas layers [51]. Unconventional oil and gas reservoirs often have low permeability, leading to borehole collapse, loss, high friction, bedding, and nano-micro pore development, posing technical challenges in the petroleum sector [52]. The solutions to these problems have been made possible by advancements in drilling fluid technology. Water-based drilling fluids (WBDFs) offer more promising futures compared to oil-based drilling fluids (OBDFs) [53]. The development of WBDF technology can be divided into four phases: initiation, development, improvement, and a second cycle of development [54]. Drilling fluid additives and technology experienced rapid development between the 1970s and 1980s, leading to improved parameters like fluid temperature resistance and salinity tolerance [48]. The need for better drilling fluid performance has arisen with the expansion of exploration and development into areas with more complicated geological conditions since the turn of the twenty-first century [24]. High-performance WBDFs with different properties have been developed, including film-forming WBDFs and low-invasion drilling fluids to enhance production and decrease drilling costs. In recent years, nanotechnology has significantly addressed challenges in the oil and gas industry, particularly in drilling engineering and drilling mud formulation [48]. Nanoparticles have been employed in drilling muds to drill elevated-temperature wells, enhance oil recovery, and improve drilling efficiency [50]. The application of nanotechnology in drilling fluids has shown promise in altering wettability, reducing interfacial tension, and improving flooding in chemically enhanced oil recovery [55,56].
Several drilling fluid technologies for formation protection have been created over more than 50 years of research to lessen the formation damage caused by drilling fluids [57,58,59]. These strategies rely on short-term plugging techniques, including shielding, physio-chemical films, and bionic technologies, progressively enhancing formation protection [60]. The fundamental idea is to temporarily block a zone with exceptionally low permeability close to the borehole wall, lessening the possibility of reservoir damage by preventing drilling fluids from entering the reservoir [61]. Following well completion, formation protection goals are achieved using perforation, dissolving, and flowback techniques. However, conventional methods do not fully address the effects of borehole collapse, loss, and high friction on formation damage, especially in unconventional oil and gas wells with horizontal drilling operations [27]. Wellbore stability technology primarily uses chemical inhibition and plugging techniques to prevent collapse and fracture. Still, these techniques have limitations and do not effectively improve borehole wall rock strength and tightness [14].

4. Advantages of Applying Nanotechnology in the Oil and Gas Industry

4.1. Enhancing Drilling Fluid Performance through the Integration of Nanoparticles

It has been discovered that adding nanoparticles to drilling fluids can considerably improve their performance and solve some of their current problems [62,63,64]. Nanomaterials have better or surprising properties compared to surface-dependent materials because of their high surface area-to-volume ratio. For a number of years, scientists have been researching the use of different nanoparticles as rheological modifiers and fluid-loss reducers in water-based muds (WBMs). They have improved rheological properties utilizing nanoparticle concentrations that are quite low (1.0 wt%). These materials’ nanoscale sizes have also been shown to be more effective than microparticles at blocking pore mouths in shale. Thinner filter cakes and less fluid invasion are the outcomes of this. Drilling fluids, also known as muds, are used during drilling operations and must have the right rheological characteristics to carry drilling cuttings from the drill bit to the well annulus and then separate them at the well surface. In addition, these fluids cool the drill bit and reduce friction between the drill string and the rock being drilled. While stabilizing the wellbore and maintaining these qualities, the fluids must also avoid permanently harming the rock formation. This is particularly crucial when drilling reservoir sections because formation damage can change permeability, resulting in a reduced rate of penetration (ROP) and a direct impact on the effectiveness of oil and gas production [65].
Different drilling fluid systems have shown the benefits of using nanoparticles in drilling fluids. Water-based muds (WBMs), oil-based muds (OBMs), and synthetic-based muds (SBMs) are a few examples of these systems. In order to reduce the drilling process’s negative effects on the environment, WBMs are frequently used for drilling into aquifer parts of wells. While SBMs use various environmentally favourable fluids such as glycol, glycerines, and glucosides, OBMs commonly use diesel as the primary fluid. Material advantages and economic benefits are considered for a particular drilling segment when choosing a drilling fluid system [66]. Nanoparticles in drilling fluids have shown promise in improving wellbore stability and filtration management. Drilling fluids can contain nano additives to act at the solid–liquid interface between the drilling fluid and the rock formation, such as nano silica and GO (graphene oxide). Nanoparticles can also help fluid rheology, which is vital for drilling oil and gas wells. This fluid characteristic may be improved using nanocellulose. In addition, several crucial features of drilling fluid design, like drill bit cooling, pollution control, and safety precautions, show potential for nanoparticle use [67].
When drilling into shale-containing strata, wellbore stability is an important factor to take into account. Drilling difficulties, including stopped pipes and hole collapse, can result from water infiltration into shale formations undermining the wellbore. When compared to commercially available shale inhibitors that are not nanoparticle-based, water-based mud (WBM) formulations using nano silicas have shown superior shale inhibition. Due to the exceptionally low permeability of shale formations, conventional filter-cake-forming procedures are not effective when drilling into shales using WBMs. When creating drilling fluids for use with shales, consideration must be given to the mineralogy of the formation since different kinds of clay found in shale formations respond to water in various ways. Nanoparticles can be made to penetrate far into the formation and block the shale pores, resolving the problems associated with drilling through water-sensitive shale formations [68]. It has been discovered that shale core samples containing 17% to 20% smectite can be efficiently plugged using high concentrations (5 to 50 wt%) of nano silica. High quantities of nano silica in the field, however, have limited economic feasibility. For drilling operations, attempts have been made to generate more active varieties of nano silica. In WBMs with 28% smectite/illite mixed layers, nano silicas functionalized with alkylamines show increased shale-inhibiting activity at concentrations below 3 wt%. These nano silicas demonstrate even greater effectiveness in preventing clay reactions in shales when subjected to seawater-based drilling fluids when paired synergistically with polymers like polyanionic cellulose.
Furthermore, compared to silicate muds known to impede shales, these nanoparticle systems require fewer caustic fluids. Silicate muds normally require a pH above 12, while nano silica drilling fluids can be adjusted to a pH below 10. In drilling operations, it is essential to increase wellbore stability by employing nanoparticles. With respect to the clay content found in shales, nano silicas have been used as agents to block pores in rock formations and lessen the negative impacts of water reactivity [69].
When drilling through varied rocks, wellbore strengthening is essential. It is equally crucial for gas and oil wells dug in sandstones, limestones, and shales. Using ferric hydroxide nanoparticles is one method to promote wellbore strengthening during drilling operations. These nanoparticles were used as a loss circulation material (LCM) in a study by Yang et al., strengthening the wellbore and reducing fluid losses. The final compressive strength values for the ferric hydroxide nanoparticle formulations tested on sandstone core samples ranging from 16.5 to 27.5 MPa were 70% greater than those for unexposed cores [70].

4.2. Harnessing Nanoparticles for Enhanced Filtration Control and Thermal Conductivity in Drilling Fluids

The use of additives is necessary to stop drilling fluid filtration into rock formations having higher permeability, such as sandstones and limestones [71]. Normally, drilling fluids are weighted to maintain a positive overbalance against the pore pressure of the formation. At the point where the fluid and rock meet, a filter cake forms to stop fluid loss from the drilling mud. The filter cake, also referred to as the “mud cake”, should immediately acquire this impermeability and be resistant to fluid loss. The mud cake must also be thin for subsequent cementing procedures in order to make it easier to remove it after drilling. Inadequate cement bonding may be hampered by leaving a thick mud cake on the formation, which could result in leaks, mechanical flaws, and cement debonding from the formation [72]. Some nanomaterials have shown synergistic interactions with clays, including bentonite, which not only reduce fluid filtration into the formation but also improve the fluid’s rheological characteristics. In this approach, ferric oxide nanoparticles in bentonite muds have produced encouraging results. It is thought that these nanoparticles intercalate the platelets of bentonite clay, displacing complexed ions (calcium and/or sodium) at high temperatures, hence enhancing filtration control. During tests at 120 °C, a loading of 0.5 wt% of nano-Fe2O3 caused a 22% reduction in filter-cake thickness and a 17% reduction in filtrate volume. Testing on the effectiveness of Fe2O3 nanoparticles with diameters ranging from 3 to 30 nm in drilling fluids at temperatures up to 200 °C revealed increased filtration control benefits, with a 28% drop in filtrate volume compared to the control without iron nanoparticles. Sandstone core samples from water-based muds (WBMs) that contain xanthan and graphene oxide (GO) have also shown value in filtration control. The filter cake that was created during the standard American Petroleum Institute (API) test for fluid-loss control was noticeably thinner in the case of WBMs with GO when compared to commercially available drilling fluids. When compared to the commercially available additive, the amount of filtrate that was collected during the test was likewise marginally less [73].
When designing drilling fluids for harsh conditions, particularly those with temperatures above 120 °C, which can cause viscofiers in the fluid to thermally degrade and have an adverse effect on the shear response, nanoparticles have demonstrated promise. Mainly, when drilling fluids contain significant levels of brine, materials that are resistant to oxidation, such as polymers and nanomaterials, are extremely helpful in such hostile settings. Instead of utilizing solid particles like barite, which can permanently plug pore throats and impair rock permeability, high brine concentrations are used to densify the fluids. As a result, these additives are normally not used in reservoir well sections where formation damage must be kept to a minimum. Instead, using soluble high-density salts such as calcium chloride, sodium formate, sodium bromide, and zinc bromide is advised in these circumstances. These fluids need yield points greater than 1 Pa in order to carry, suspend, and convey the drilling cuttings efficiently. Conventional additives like guar gum, xanthan, and crosslinked starch are frequently utilized to reach these yield points. However, at temperatures exceeding 120 °C in saline settings, oxidative degradation to these compounds becomes a severe issue [74].
In brine-based drilling fluids, cellulose nanofibers (CNFs) have shown promise as rheology modifiers. According to research by Heggset et al., functionalizing nanocellulose through controlled oxidation with the 2,2,6,6-tetramethylpiperidine-1-oxyl radical (TEMPO) improves the temperature tolerance of water-based muds (WBMs) made from caesium formate and sodium formate brines when compared to non-functionalized nanocellulose and other common polymers like guar gum and xanthan. CNFs and cellulose nanocrystals (CNCs) have both demonstrated the capacity to raise yield points in bentonite-containing WBMs. It is thought that CNCs interact synergistically with polyanionic cellulose to boost yield points and fluid thixotropy by forming complexes with bentonite platelets. Further advancements in filtration control and rheology are seen when cationic CNCs (caCNCs) and carboxylate-functionalized CNCs (cCNCs) are utilized in low-solid WBMs. Although CNFs and CNCs have the potential to change their rheological properties, they perform very poorly in terms of filtration control when compared to other nanoparticles like iron oxide nanoparticles, graphene oxide (GO), and nano silica [75].
Electrical and thermal conductivity are essential factors in drilling operations. It is necessary to use a thermally conductive drilling fluid to effectively remove heat from the drill bit. Similar to resistivity logging, a method frequently employed for seeing the wellbore, an electrically conductive drilling fluid is desired. Investigations have focused on how nanoparticles affect the thermal and electrical characteristics of drilling fluids. The high surface area-to-volume ratios of nanomaterials are thought to play a significant role in improved electron conduction and increased heat transfer rates. In polyanionic cellulose water-based muds (WBMs), studies have shown improvements in thermal conductivity using nanosized tin oxide and 50 nm copper/zinc oxide, with increases in thermal conductivity of up to 53% at concentrations of 1 to 5 × 10−3. The 50 nm zinc oxide nanoparticles have also been found to improve electrical conductivity. Drilling activities are highly unsafe due to hydrogen sulphide (H2S) discharge from specific rock formations. Field workers must be careful around H2S because exposure to levels as low as 600 ppm can quickly result in death. Additionally, H2S severely corrodes well-digging machinery. Iron oxide (Fe3O4) and zinc compounds like zinc oxide (ZnO) and zinc carbonate are frequently used to scavenge H2S. In comparison to bulk zinc oxide, nanosized zinc oxide particles have more favourable kinetics when reacting with H2S due to their high surface area-to-volume ratio. At room temperature, nanoparticle samples have shown the ability to remove 800 ppm of H2S from mud entirely in 15 min, outperforming bulk ZnO, which can only remove 2.5% of H2S in 90 min [22].

4.3. Nanoparticle Advancements in Drilling Fluids: Optimizing Filtration, Rheology, and Environmental Impact

In commercial drilling, managing drilling fluid waste is a major environmental concern, and actions are being taken to mitigate the effect of drilling activities utilizing non-aqueous fluids on the metal content of marine sediments. It is common knowledge that drilling mud pits at drilling rigs can contain sizable concentrations of hazardous heavy metals [76]. Groundwater penetration and soil pollution are risks associated with uncontained heavy metals. In order to address this problem, it has been demonstrated that functionalized nanomagnetic Fe3O4 particles are efficient at removing dangerous pollutants like cadmium from both water-based muds (WBMs) and oil-based muds (OBMs). These nanoparticles function as magnetic nano adsorbents that can draw out cadmium from drilling fluids in a targeted manner. These nanoparticles have nano-Fe3O4 in their centre, which is covered with polyacrylic acid chains. These chains have amino groups functionalized at their ends and are covalently joined to the iron oxide core. A WBM with 5 ppm of cadmium is treated with these nanoparticles, and 90% of the cadmium is quickly eliminated from the mud. The performance of drilling fluids has been improved in field tests using nanoparticle additions, which are now being conducted. However, the price of nanoparticles continues to be an important factor, and lowering it will be essential for their widespread use and commercialization in the industry [3].
The impact of zinc oxide (ZnO) nanoparticle sizes on the rheological characteristics of water-based mud (WBM) is examined over a temperature range that includes room temperature and downhole conditions (roughly 90 °C) in the study by Alkalbani et al., 2022 [77]. Adding nanoparticles—specifically, zinc oxide (ZnO)—to drilling fluids is a viable way to improve rheological characteristics, filtering efficiency, and hydrogen content. There are three distinct steps to the experimental design. The first stage determines the standard WBM’s baseline rheological characteristics. Two sizes of ZnO nanoparticles (10–30 nm and 30–45 nm) are added in later stages to assess the rheological characteristics at different temperatures. The results showed that adding ZnO nanoparticles significantly improved the rheological characteristics of WBM by 40% to 65%. The bigger nanoparticle size (30–45 nm) further boosted fluid stability under a variety of situations. The research highlights the need to address the environmental effects and safety issues associated with these nanoparticles in addition to helping to clarify the complex relationship between nanoparticle size and rheological properties. The study contributes to the ways in which the sizes of ZnO nanoparticles influence the rheological characteristics of WBM, laying the groundwork for further investigation and possible uses in the oil and gas industry [77]. If the whole impact of the findings on the environment is considered, the results could serve as a stimulus for the prudent integration of nanoparticles in downhole settings.
In a further study by Alkabani et al., 2022 [78], the research explores the use of nanotechnology in drilling fluids with the goal of improving filtration rates, hydrogen concentration, and rheological characteristics. This exploration is especially relevant in light of the use of sophisticated and specialized drilling and production techniques on a variety of hydrocarbon reservoirs, including shale gas and oil as well as deep-water reservoirs. The justification for using nanotechnology is its capacity to provide an economical and ecologically responsible substitute for traditional sophisticated methods. Water-based mud’s rheological properties are thoroughly investigated by means of an experimental design using a response surface method that takes temperature variations and nanoparticle concentration into account. The experimental setup includes a rotating viscometer for the accurate measurement of essential parameters, including viscosity, yield point, and gel strength. In addition, the fluid sample is subjected to a variety of temperature settings using a consistometer, which simulates the spectrum up to downhole temperatures (90 °C). The incorporation of zinc oxide nanoparticles into water-based mud significantly improved its rheological characteristics, according to empirical studies. In particular, there is a 50% increase in viscosity and gel strength and a notable 80% increase in the yield point. In addition, concentrations of zinc oxide nanoparticles greater than 1 wt% (6 g) and lower temperatures (below 40 °C) appear to produce the best and most effective results in terms of improved rheological characteristics.
The study by Alkalbani et al., 2023 [79], aimed to enhance the rapidly developing subject of nanotechnology in drilling fluids by studying the rheological characteristics of water-based mud (WBM) that contains silica nanoparticles through an experimental approach, mainly for deep-well applications. Investigating nanoparticles—especially silica dioxide (SiO2) nanoparticles—is an attempt to tackle the relatively new use of nanoparticles in the drilling industry. The study aimed to optimize the drilling process, especially in formations that present problems, by adding nanoparticles to standard drilling mud as external additives or doping agents. The effects of adding silica dioxide nanoparticles to a standard water-based mud at weight concentrations between 0.86 and 2.57 wt% on its rheological characteristics were investigated. These nanoparticles are mixed with a WBM including calcium carbonate (CaCO3), xanthan gum, sodium chloride (NaCl), starch, HT starch, and caustic soda (NaOH) to create the nano drilling mud. This nanofluid is designed to be used in deep and intricate wells, and its rheological behaviour is carefully evaluated. The fluid samples are heated to high temperatures in a roller oven that has a temperature range of 38 °C to 232.2 °C, which simulates downhole conditions. It was observed that at 0.86 wt%, silica dioxide nanoparticles exhibit a viscosity of 33 cP and a gel strength of 3–5 lb/100 ft2, while at 2.57 wt%, the values are 44 cP and 4–5 lb/100 ft2, respectively. Significantly, at 1.71 wt%, the ideal concentration shows much more improvement, reaching 55 cP and 4–5 lb/100 ft2. A comparative study shows that 1.71 wt% concentration of silica nanoparticles in WBM performs better than 0.86 and 2.57 wt% concentrations.

4.4. Nano Silica Modernizing Oil Well Cementing: Enhancing Strength, Reducing Porosity, and Mitigating Gas Migration Risks

Cement set accelerators are essential in the cementing of oil wells by speeding up the transition from the liquid (slurry) phase to the solid phase [80]. The expenditures associated with the waiting on cement (WOC) phase can be reduced through careful cement design. Accelerators are also utilized to reduce gas migration to achieve good zonal isolation. Commonly used as accelerators are inorganic salts like calcium chloride; however, they can also increase cement porosity, which is not ideal. Nano silica has been suggested as a possible substitute cement addition to overcome this problem. In comparison to conventional accelerators, nano silica demonstrates a number of benefits, including a decrease in porosity and an increase in mechanical strength of up to 35%. Nano silica is a useful accelerator and reinforcing nanomaterial due to its high surface area-to-volume ratio and its impact on the formation, growth, and seeding of C—S—H (calcium silicate hydrate) in cement. Due in part to the acceleration of the pozzolan reaction between hydrated lime (Ca(OH)2) and SiO2, nano silica speeds up the production of C—S—H. Nano silica hence has the potential to be used in a variety of oil and gas well cementing scenarios, such as cementing in lower temperature zones. Nano silica also addresses cement’s problems with gas migration. Following cement placement in the well, gas migration is possible, and hydration processes reduce cement’s gel strength. The hydrostatic pressure that the cement slurry density was exerting on the nearby rock formation is eliminated by this loss of gel strength. As a result, channelling occurs before the cement has the necessary mechanical strength to withstand fluid/gas ingress. The cement also becomes more susceptible to pore pressures from the formation. By reducing the period between the cement’s weak gel state and its solidification into a structurally robust material, nano silica and C—S—H nuclei have shown their capacity to lower the danger of gas migration [55].
Thus far, the most successful application of nanotechnology in drilling has been the creation of drilling fluids that strengthen wellbore integrity and improve their rheological properties. Another noteworthy development is strengthening the durability and resilience of materials used in drill bits, casing, and related gear. Additionally, there are succinct summaries of developments in ball packers that produce multiple fractures and improve cement integrity [49].
Because shale swells as water seeps into its pores, drilling in shale zones with inexpensive, water-based fluids can create problems for the integrity of the wellbore. Drilling fluid water is drawn into shale pores by the substantial osmotic pressure caused by the brine’s higher ionic strength than that of the drilling fluid water [81]. For rocks with nano-sized holes, such as shale, where the osmotic force outweighs physical obstruction, this technique is insufficient even though colloidal particles in the drilling fluid create a filter cake to prevent water entry [82]. In an attempt to tackle this problem, silica nanoparticles were added to the drilling fluid as reported in a study by Hussein et al., 2015 [83]. These studies successfully demonstrated how well nanoparticles prevented water ingress (see Figure 4). The physical obstruction of far smaller pore cross-sections by trapped nanoparticles is the mechanism underlying this prevention. While this approach slows down low-ionic-strength water seepage into shale pores, efforts are still being made to create more effective nanoparticle surface coatings in order to completely eradicate the osmotic pressure differential. This method was successfully field-tested by MI-Swaco and Baker-Hughes [18].
Figure 4 shows the Shale Membrane Tester (SMT) test apparatus and the change in the pressure in shale when silica nanoparticles are injected using a water-based drilling fluid. Even after flushing with brine for 15 h, the pressure and permeability in the shale sample was continuously low, suggesting that silica nanoparticles were physically blocking the shale [85]. Using functional nanoparticles as emulsion stabilizers for drilling fluid emulsions could be one way to improve drilling fluids. Oil-based mud is used to stop water incursion from causing shale formation swelling. Completion issues are presented by the filter-cake, drilled solids, and oil-coated borehole [86]. Reversible drilling-fluid emulsions that are water-external for completions and oil-external for drilling have been developed and used successfully to address this. Drilling and completion efficiency could be greatly increased by using functional nanoparticles as emulsion stabilizers, which could produce either water-in-oil or oil-in-water emulsions based on drilling and borehole conditions or external control [87,88]. The stability of such emulsions may be controlled by employing SPM-NPs as an emulsion stabilizer and an external magnetic field.
Drilling fluid loss, rheological changes, lost circulation, and wellbore instability are frequent problems in the context of drilling operations that must operate in challenging environments such as deepwater, high-temperature, high-pressure (HTHP), and unconventional situations [89,90,91]. By lowering fluid loss, preserving optimal rheological properties, and boosting wellbore strength [92], the addition of NPs, such as silica and iron oxide NPs, to drilling fluids has demonstrated promising results [68,93,94]. According to Okoro et al., 2020 [95], NPs are assumed to reduce fluid filtering by filling in spaces between larger particles to generate a thinner, non-erodible, and low-permeability filter cake. This process results in a less permeable seal.
Because of their minuscule grain sizes, alloys with nano-scale microstructures have exceptional mechanical characteristics. Drilling could undergo a major revolution if lightweight, high-strength steel that is affordable and mass-produced is developed for tubing and casing [96]. While much research has been carried out on producing aluminium alloys [44] and nano-ceramic coatings [97], relatively little research has been carried out on steel, even though steel is used widely and has practical importance.
The DOE Idaho National Engineering and Environmental Laboratory came up with the brilliant steel breakthrough NanoSteel Co., licensed to NanoSteel Co. in Maitland, Florida [98]. Their proprietary technique yields this Super Hard Steel (SHS), which combines high strength, toughness, corrosion resistance, and hardness. Its homogeneous distribution of fine borocarbide and refined nanoscale microstructure are the sources of these characteristics. When applied as a coating, the SHS alloys effectively create metallic glasses that, when heated, become a nanoscale composite coating.
Another possible use for nanoparticles is covering the inside or outside of tubing to (i) lessen wear and corrosion; (ii) increase smoothness to lessen pressure drops; and (iii) reduce paraffin or scale deposition. The use of smart coatings, or structured layers of specific alloy nanoparticles, to prevent corrosion and abrasion, has been researched. Chemical mechanical polishing, also known as chemical mechanical planarization or CMP, is a well-proven technique for producing remarkably smooth solid surfaces. It is frequently employed in the production of semiconductor wafers and related substrates. This method creates incredibly clean solid surfaces by using a polishing pad in combination with an abrasive and corrosive nanoparticle slurry. Notably, CMP is currently among the biggest users of nanoparticles [99]. The NSF/SRC Engineering Research Center for Environmentally Benign Semiconductor Manufacturing at the University of Arizona is a great source for a thorough guide on CMP [100]. Enhancing hardbanding materials, like those created for casing, is another possible use of nanoparticles in tubing. In order to increase abrasion resistance and enhance sealing capabilities, research is concentrated on developing particular nanoparticle dispersions that may be applied to tool joints and other tubing surfaces [101].

4.5. Innovations in Cement Technology: From Self-Healing Additives to Nanosensors, Paving the Way for Durable and Enhanced Well Integrity

Using a cement viscosity enhancer, using cement-set accelerators, regulating cement density, embedding nanosensors in cement, incorporating nanoparticle additives that promote microcrack self-healing, and controlling cement density are noteworthy research endeavours aimed at improving cement integrity. In deepwater well completion, controlling the cement-set time becomes critical because of the large cement-filled casing-formation gap and the chilly seawater. Conventional cement-set accelerators, such as inorganic salts like calcium chloride (CaCl2), are known to increase cement permeability while shortening the set time. According to Thibaud et al., 2018 [102], silica nanoparticles speed up cement hydration, especially at low temperatures (around 15 °C). When applied to fresh cement, nano silica with a specific surface area of between 60 and 650 m2/g can improve its rheological characteristics, which can lead to improved concrete consistency (self-compacting concrete, for example) and decreased bleeding and segregation [103].
A cement viscosity enhancer is used to reduce the amount of sulphate and chloride ions that diffuse into the cement, weakening its structure and perhaps causing cracks. Earlier attempts to increase cement durability concentrated on producing denser, less porous cement prone to cracking. The National Institute of Standards and Technology used a revolutionary technique that effectively doubled the cement’s service life. They achieved this by adding nanoparticles to the solution, increasing its micro-pore viscosity and preventing ion migration. It has been discussed that incorporating cellulose nanofibers (CNFs) and nanocrystals (CNCs) into cement paste increases the flexural strength of cements [104].
The idea behind self-healing additives is to incorporate dispersed nanoparticle catalyst and polymeric material into microcapsules that are embedded in poured cement. When a microcapsule and a catalytic particle come into contact, a polymerized healing agent that seals the microcrack is created [105]. It is possible to record required data by integrating nanosensors into cement; however, active data transmission is still difficult because the present technology does not provide a power supply. An externally powered detector could access data contained in passive nanosensor particles [106]. While active data transmission is still a technological problem, the successful embedding of microsensors into structural concrete has been accomplished [107]. An innovative technique for causing fractures has emerged as a result of the extensive use of repeated fracturing in horizontal wells. With this method, the horizontal wellbore’s sleeves are activated one at a time, utilizing ball seat technology. This novel technique uses nano-composite balls with significant strength to perform their function during fracturing, in contrast to traditional balls and ball seats, which must be removed after fracturing.
Controlled electrolytic metallic (CEM) materials are used to apply the novel ball seat technology and its related counterparts effectively. These materials have a metallic and/or ceramic nanostructured covering designed for regulated, adaptable in situ deterioration in subterranean settings [108]. Because of their high strength and low deformation, the CEM materials make staged hydraulic fracturing possible by allowing for localized pressure buildup. By using dissimilar metal composites to build micro-galvanic cells, lightweight, highly durable materials with controlled corrosion rates and the ability to break down in situ were created [109]. Because of the engineering of the micro/nano-matrix and the incorporation of secondary nanoscale metallic and/or ceramic improvements, they possess ultra-high strength. High-strength disintegrating materials (HSDMs), which have a higher rate of disintegration, greater strength, and decreased weight, were also invented by Ejileugha et al., 2022 [110]. These materials are made up of powdered metallic alloys reinforced with ceramic or metallic coatings.
High viscosity is a crucial criterion for fracturing fluids during the fracturing stage, as it minimizes fluid seepage into adjacent matrix zones and creates a significant pressure differential in the propagating crack [111]. But after the proppants are in place in the fracture, it is crucial to remove the fracturing fluid carefully so as not to obstruct the flow of gas, water, or oil that follows. Attaining these two goals at the same time is difficult, but there is a way to manage fracturing fluid rheology externally by combining polymers and nanoparticles [112] in the following way. High viscosity is produced during fracture when polymer molecules are joined to functional nanoparticles in the form of dendrimers or networks. The fluid viscosity decreases during the removal phase when the polymer molecules separate from the nanoparticles, making removing the fluid from the freshly formed crack easier. Ferrofluid has also been proposed as a fracturing fluid for external control. One major issue with today’s shale oil and gas operations is the large amount of water used in hydraulic fracturing. To address this, CO2 foam has been used as a fracturing fluid. Ikram et al., 2022 [113], were able to lower the water volume fraction in CO2 foam to as low as 2% while keeping the foam’s apparent viscosity at about 100 cp during flow in bead-packs by utilizing the strong stability of CO2 foams produced with silica nanoparticles. According to model simulation results reported by Jia et al., 2022 [114], using such foams would result in larger fracture development and noticeably faster cleanup times. To prevent fines from migrating into the area close to the wellbore, fracture proppants coated with nanoparticles have been produced [115]. Through surface adsorption and charge interaction, nanoparticles added to viscoelastic surfactants (VESs) created a more resilient dynamic network of VES micelles. By maintaining fluid viscosity at high temperatures and preventing VES micelle breakdown, this strengthened network decreased fluid leak-off. Because of their small size, nanoparticles readily flowed back with the generating fluid after internal breakers were added to disrupt VES micelles [116].
In order to create secure sealing parts for packers, joints, o-rings, and other well completion system equipment, elastomers are essential. Because these finishing systems operate in harsh environments with elevated temperatures and/or pressures, the durability and improved elasticity of the elastomers utilized are critical. Field tests of these materials are assumed to have been carried out. Polymers integrated with nanomaterials, such as CNT, have been created and demonstrate noticeably enhanced elasticity compared to conventional materials [117].

4.6. Nanoparticle Innovations Transforming Oil Industry Practices: From High-Temperature Insulating Packer Fluids to Precision Conformance Control

Deep oil wells and steam injectors have difficulties due to uncontrolled heat loss, which can result in problems including clogging, the deposition of paraffin and asphaltenes, and poor steam quality. High-viscosity high-temperature insulating packer fluids (HTIPFs) with nanomaterials are being studied by Halliburton [118]. In order to accomplish desirable qualities, including low convection, high thermal isolation, high viscosity, high strength gelling, and high temperature stability, this packer fluid uses an intermolecular association synthetic macromolecule system.
Coatings now have multiple functions thanks to the use of functional nanoparticles on solid surfaces: (i) reducing surface abrasion; (ii) reducing corrosion and scaling on facility surfaces inside and outside; (iii) exhibiting the capacity to self-repair minor surface damages; and (iv) identifying emerging corrosion beneath the coating to provide early warnings. The novel surface treatments created by C3 International are particularly noteworthy for toughening solid surfaces to lessen abrasion [119]. Their treatment entails the use of 3 nm crystallites to encapsulate the thin oxide film in the grain boundaries of the surface. This makes the surface treated with nanoparticles extremely resistant to wear, extending its lifespan and minimizing downtime. This process, known as “molecular implantation surface treatment (MIST)”, has undergone extensive testing by researchers at the High-Temperature Materials Laboratory at Oak Ridge National Laboratory, one of the notable producers of nano-structured coating materials which produces nanoscale-grain coating materials with its NanovateTM technology [120]. Their high-permeability surface coating for low-frequency magnetic shielding is called Nanovate-EM, and their enhanced erosion and impact protection for composite parts is called Nanovate-NS metal coating. These coatings are designed to minimize friction and maximize flow while reducing corrosion, wear, and erosion, especially for oil field tubular goods [121].
The generated oil may form highly stable water-in-oil emulsions because it contains natural surfactants or asphaltenes. It is a costly procedure to separate this emulsified water from the oil by causing the emulsions to destabilize and coalesce. Likewise, extremely stable oil-in-water emulsions can occasionally form in the created water, necessitating expensive oil droplet removal. The use of functional nanoparticles that easily adsorb to the oil/water interface and break apart emulsions has been the subject of extensive investigation. By first adhering to the oil–water interface and then extracting the MCNTs connected to the oil droplets from the water by means of a magnetic gradient, magnetic carbon nanotubes (MCNTs) are able to extract oil droplets from generated water [122]. In another example, surface-active ethyl cellulose grafted magnetic nanoparticles (M-EC) were used to remove water droplets from diluted bitumen emulsion, removing over 90% of the water. Ten times over, the utilized M-EC was regenerated and reused while keeping the water droplet removal effectiveness constant [123]. As shown in Figure 5, surface-engineered SPM-NPs with an opposing surface charge that sticks to the droplets effectively separated negatively charged oil droplets from generated water by as much as 99.9%. Quantitative descriptions of the dynamics involved in the magnetic separation of SPM-NPs from water were obtained using model simulations [124]. The method of eliminating oil droplets with surface-engineered SPM-NPs that bind to the droplets countersurfacely is shown in Figure 5. An extra picture at the top shows how clean water is separated from SPM-NPs that are affixed to oil droplets.
A relatively new membrane process typically removes multivalent cations and submicron-size natural organic matter [126]. Ceramic NF membranes are becoming more common because of their high chemical, mechanical, and thermal stability. Although they are susceptible to fouling by waxes and asphaltenes, they are efficient in the high flux separation of oils, emulsions, and silts. Nanofiltration (NF) is a filtration method that falls between traditional filtration, which typically removes colloidal particles, and reverse osmosis (RO), which can remove monovalent ions. As a pre-processing stage before RO, NF is frequently utilized in conjunction with an RO process to desalinate water [127]. The second use is the extraction of asphaltenes from natural gas condensates; the value of the condensate is reduced in the presence of large-size asphaltene aggregates. In this case, low-cost eradication with NF is regarded as an economically appealing approach. Current applications of nanofiltration (NF) in the oil industry include three primary domains. The first use is purifying water extracted from oil and gas reservoirs to remove organic materials and colloidal particles. Reusing produced water for procedures like chemical injection or waterflooding is made possible by this treatment, which also guarantees safe disposal into subterranean formations or at the ground’s surface. Thirdly, the removal of H2S and/or CO2 from a natural gas stream via NF is a topic of research for several investigators. There are several suppliers of NF supplies and equipment on the internet.
Wax or asphaltene particles frequently precipitate out of the oil during crude oil extraction from deep-sea offshore fields due to the extreme temperature differential between the normally hot reservoir and the near-freezing conditions at the seabed [128]. Significant delays to oil production operations are caused by these particles, which tend to deposit on the inner surfaces of oil transport pipelines [129]. Methane hydrate deposition in natural gas pipes also poses a significant obstacle. Electrical resistive heating is usually applied using a “pipe-in-pipe” arrangement to create an insulated gap or a heating jacket wrapped around the pipe. SPM-NPs can effectively produce localized heat when exposed to an oscillating magnetic field in the range of roughly 500–1000 MHz. After specifically surface-coated SPM-NPs are injected into a patient’s circulation, allowing them to adhere to the cancer cells, a heating method known as “hyperthermia” [130] is used to eradicate the cancer cells. Numerous studies have been carried out in this field.
A coating placed to the inside surfaces of subsea oil pipelines may have embedded SPM-NPs, an idea that Mohanty et al., 2022 [131], investigated because of the possibility of highly localized and energy-efficient hyperthermia heating. An oscillating magnetic field can be used to create a thin, heated layer of wax (or asphaltene, hydrate) at the interface of the “nano-paint” with the inner pipe surface. As shown in Figure 6, this heat causes the deposits to separate from the pipe walls and allows them to be taken downstream by the flowing oil. Razali et al., 2022 [121], suggested utilizing a tiny, battery-powered magnetic oscillation emitter that floats freely and moves with the oil—also known as a “intelligent pig”—to maintain regular oil transport operations. The researchers assessed the heat-generating capacity of several SPM-NPs in various dispersing media and conducted preliminary simulations to determine the practical viability of using an intelligent pig.
Delivering acid to the intended reservoir zone while avoiding harm to surface facilities and wellbore equipment is a major difficulty for matrix acidizing. A strong acid is needed to form high-permeability flow channels in the matrix zones, but this acid usually reacts with the metals in the wellbore and surface facilities before it gets to the reservoir area. A possible solution to this problem is to use double emulsions as the stimulating (acidizing) fluid, particularly an acid-in-oil-in-water emulsion. Using this method, the acid is kept out of the outermost water phase that comes into contact with the wellbore hardware and surface facilities. It is usually difficult to generate such double emulsions using surfactants. Nonetheless, nanoparticles can be used to create stable emulsions with almost uniform droplet sizes (like acid-in-oil), and by emulsifying this in water, another kind of nanoparticle or a surfactant can be used as a stabilizer [133]. These double emulsions have the potential to be very potent well-stimulation agents if they can be produced with very little oil and maintain their stability until they reach the deep reservoir zones that are the focus. One such technique for localizing chemical delivery is microencapsulation in a polymer shell. Since microfluidic techniques have made it possible to reduce capsule sizes to the sub-micron range, this technique, which was initially designed for the targeted and progressive release of pharmaceuticals, holds potential for subsurface applications [134]. These methods entail spraying a stream of emulsion droplets from a nozzle that are distributed across a liquid that can form a polymer film. Through the manipulation of fluid viscosities, flow rates, and nozzle diameters, a thin liquid film is formed around each droplet, which then solidifies.
It is possible to envision three different uses for nanoparticles that will increase the efficacy of conformity control: enabling the blocking and opening of particular reservoir zones in response to fluid injections, and doing away with the requirement for pricey and unstable downhole hardware. The idea of developing an “on-command” check valve fluid without requiring complex wellbore hardware is based on ferrofluid, which is used as an actuator medium in the automobile sector [135]. Similarly, it is possible to manage crosslinked polymers or emulsions stabilized using paramagnetic nanoparticles to act as an externally controlled fluid that changes its rheology in reaction to a magnetic field. Applying a magnetic field locally along the wellbore allows precise injection or extraction from particular zones. Other uses, such as enhancing enhanced oil recovery (EOR) fluids in challenging reservoir conditions, can also benefit from this strategy, enhancing compliance control fluid stability in difficult reservoir environments. In these difficult circumstances, nanoparticle-stabilized emulsions or foams provide significant advantages for EOR fluids. Another cutting-edge technique is using thermo-sensitive polymer-grafted SPM-NPs to create a gel that solidifies in response to temperature changes. Gel conformance control can be tackled by introducing this polymer into a selected reservoir zone and heating it selectively using magnetic hyperthermia. This method keeps passageways in low-permeability zones open while high-permeability zones are blocked. In a study on this conformance control technique, Singh et al., 2022 [80], tested the effects of mixing four distinct gel-forming polymers with SPM-NPs: curdlan, methyl cellulose (MC), hydroxypropyl methylcellulose (HPMC), and hydrolyzed polyacrylamide-polyethylenimine (HPAM-PEI). The investigation revealed gel formation brought on by oscillations in the magnetic field.

5. Nanotechnology and Environmental Impacts

Nanotechnology has become a ground-breaking applied technology in the past few decades. This development enables atomic-level manipulation, leading to novel structures with unique properties. Applications of nanotechnology are found in a wide range of industries, including engineering, medicine, and life sciences. One crucial area in which nanotechnology can have a significant impact is environmental stewardship, which addresses issues such as pollution, resource efficiency, and renewable energy production. Nanosensors are essential for accurately tracking how human activity affects the environment and for enabling early intervention in the form of care, treatment, and prevention. Ultimately, nanotechnology helps to minimize environmental harm by improving resource usage and refining current contaminants. Moreover, nanomaterials facilitate the transition from fossil fuels to renewable energy sources by improving the efficiency of clean energy generation. Businesses that use nanotechnology have the ability to produce biodegradable, environmentally friendly materials, which could help reduce pollution. This section explores the potential uses of nanotechnology to tackle major environmental issues, such as air pollution, water scarcity, municipal solid waste management, and nanomaterial safety [136].
The impact of various additives obtained from waste on attaining effective rheological qualities in drilling fluids has been brought to light by recent studies. For example, Joshi et al., 2012 [137], investigated tamarind kernel powder as a substitute ingredient for drilling fluids. The study examined how adding tamarind kernel powder affected the density of the mud. Mud density is essential for drilling fluids because it affects formation pressure management and wellbore stability. The results of the investigation showed that adding tamarind seed powder to the bentonite mixture increased the density of the mud sample. An increase in tamarind seed powder concentration produced a thicker mud sample and higher mud density overall. The samples’ observed mud densities ranged from 8.22 to 8.97 parts per thousand, deemed appropriate for adding to drilling fluid formulations. Murtaza et al., 2021 [138], also demonstrated the use of okra, an eco-friendly vegetable, as a workable replacement ingredient in drilling fluids. The effectiveness of okra as an addition was evaluated in drilling fluids with and without clay. Comparatively speaking, adding okra to drilling fluids made of clay showed a more significant improvement in rheological characteristics than adding it to fluids made of non-clay. When 2 or 3 g of okra was added to clay-based drilling fluids, the plastic viscosity (PV) increased by more than 100%, exceeding the 45.7% rise that was seen when 2 g of starch was added. Drilling fluid yield point increased because of the increase in okra content. Starch was found to be more successful in raising the yield point compared to okra. Variations in decreases were observed for each concentration when fluid loss was assessed at different levels. In addition, as Figure 7a–c illustrate, the inclusion of okra resulted in a drop in filter cake thickness, with more noticeable decreases seen at higher concentrations [139].

5.1. Nanotechnology and Waste Management

Solid waste materials encompass substances generated through human activities to fulfil their needs and subsequently introduced into the environment. The escalating population and shifts in lifestyle have rendered the production of these materials a significant contemporary environmental concern. Industrial and urban waste contribute hazardous organic and inorganic pollutants to water, soil, and air. Conventional technologies struggle to effectively eliminate these pollutants. Hence, the adoption of modern technologies, such as nanotechnology, emerges as a crucial solution to address this issue. Nanomaterials, including nanofilters, nanosensors, nano photocatalysts, and nanoparticles, play a pivotal role in waste management, offering innovative approaches to tackle environmental challenges [140].
Similarly, Ghaderi et al., 2014 [141], proposed using sustainable saffron purple petals (SPPs) as a greener substitute for drilling fluid additives. Adding SPP powder significantly increased plastic viscosity (PV) values in drilling mud. There was a commensurate rise in the PV value with increasing SPP powder concentration. In addition, the yield point of the drilling mud was noticeably higher with the addition of SPP powder than with the base mud. The incorporation of SPP powder into drilling mud demonstrated remarkable results for filtrate loss, demonstrating a progressive decrease in filtrate volume as the concentration of SPP powder increased. Food waste effectively replaces the environmentally hazardous products currently used in the drilling fluid sector. In this line, it is necessary to solve the issue by investigating the rheological properties of the additives, ensuring that it is feasible and cost-effective, and promoting the “waste to wealth” concept by researching the possibility of employing unneeded waste derivatives as additives in drilling fluids [142].

5.1.1. Nanofilters in Waste Management

Common biological and physicochemical methods are unable to remove all of the toxic and nonbiodegradable materials found in wastes and their leachate, including heavy metals, ammonia, xenobiotic organic compounds, arsenic, and inorganic macrocomponents. Pretreatment, post-treatment, and contemporary technologies are being explored to remove pollutants from the environment to address this problem. Nanofiltration can treat waste without using extra chemicals, making it a viable technique. By intensifying and concentrating the trash, this method lowers the cost of disposal and transportation. It has been shown that using nanofilters may eliminate 60–70% of COD and 50% of ammonium from the leachate. The removal of pathogens, anions, cations, arsenic, uranium, and chromium from wastewater is made possible using nanofilters. Membrane fouling hinders the nanofiltration technique, which necessitates costly cleaning or replacement. It is more economical to use materials like fullerenes to prevent biological fouling. Fullerenes function as membrane antifouling agents by inhibiting bacterial attachment and microorganism growth, which lessens the formation of organic films that can block pipes and membranes. Moreover, fullerenes prevent membrane and pipe obstructions by acting as clot-busting agents. It seems that applying these nanoparticles to membranes and pipelines is a feasible approach to lessen biological fouling [143].

5.1.2. Nano Photocatalysts and Porous Nano Catalysts

A photocatalyst is a material that, without changing itself, catalyses a chemical reaction in the presence of sunlight. When exposed to UV light, oxygen, water, and titanium dioxide produce free radicals, almost all the properties of a perfect photocatalyst. These radicals efficiently convert poisonous substances into less dangerous carbon molecules. Titanium dioxide nanoparticles have better photocatalytic capabilities than bigger particles due to their increased surface-to-volume ratio. TiO2 is widely used to treat leachate from waste material landfills. It is used in coating fixed membranes, nanocrystalline microspheres, and membranes mixed with silica. Because of its hydrophilic nature, heavy metals from wastewater can be absorbed and organic pollutants can break down. Gasification is a process that turns trash into ethanol, and porous nano catalysts are part of this process. Carbon compounds are transformed into syngas, which is mostly made up of hydrogen and carbon monoxide, under high pressure and temperature conditions in a controlled setting. Ethanol production can occur more easily when porous nano catalysts better absorb carbon monoxide molecules [144].

6. Utilizing Nanomaterials in the Drilling Operations of Deep Geothermal Reservoirs

Potential thermal energy reservoirs that might one day be commercially feasible are represented by geothermal resources. One important feature of porous rocks is the heat produced by the concentration of energy in the top crust. The densest regions are usually found close to plate borders, where a detectable geothermal gradient is common. This gradient acts as a first sign of higher temperatures than the mean value in the rocks beneath the surface. This event could be caused by a localized heat source, such as magma intrusion that occurs within a few kilometres of the crust and registers temperatures between 600 and 1000 °C. Notably, areas unaffected by recent shallow magmatic intrusions can also see the development of geothermal fields [145].
Certain circumstances, including the weakening of the continental crust, may be associated with an abnormally elevated heat flow. This would cause the crust–mantle boundary to rise and cause temperatures to rise at shallower depths. But a successful geothermal resource requires more than simply a thermal anomaly; it also needs a reservoir, or a sufficient volume of porous rock at a particular depth that can be drilled into. Hydrothermal structures include routes that allow cool subsurface water to enter reservoirs, recharging areas, and reservoirs themselves. Magmatic deposits present an economical way to produce power from geothermally accessible energy because they are positioned high enough in the crust to cause groundwater convection. Due to the rising energy needs and environmental problems related to the oil sector, geothermal reservoirs have become a promising renewable and environmentally acceptable energy source. In order to gain access to these reservoirs, geothermal exploration and drilling initiatives have significantly increased during the past few decades [146,147,148]. Geothermal energy is produced by drilling wells into these reservoirs and using circulating fluids that are either injected into or taken from the geothermal reservoirs to capture the heat of the Earth [149]. Notably, drilling methods used in oil and geothermal wells are comparable [150]. As a result, geothermal well drilling techniques are developed based on improvements in drilling technology within the oil industry [151].
Additionally, because geothermal and high-pressure, high-temperature (HPHT) oil and gas wells experience similar downhole conditions, engineers can draw comparisons and use what they have learned from drilling HPHT oil and gas wells to understand better the complexities of drilling geothermal wells. The nature of the fluids produced and the conditions in which these fluids exist are the main differences between geothermal and HPHT oil and gas wells. Geothermal wells can have temperatures in one of three ranges: low (below 150 °C), medium (between 150 and 200 °C), or high (over 200 °C) [152]. However, temperatures in geothermal wells can rise over the water’s critical point, complicating drilling and completion procedures.
The high cost of drilling operations is another issue facing the geothermal business [153]. Depending on the complexity of the geothermal reservoirs, drilling operations often account for 40–60% of the overall cost of a geothermal project, including confirmation and development drilling. The geothermal business is further burdened by these high drilling costs and lower revenues, underscoring the requirement for efficient drilling techniques [154]. Geothermal projects can become more financially viable by adopting automated operations and transferring technologies and breakthroughs from the oil and gas industry to the sector [155]. Additionally, investing in the creation of cutting-edge instruments appropriate for abrasive and hard rocks can enable faster and more effective drilling [156]. Another issue for the geothermal business is the lack of drilling and operating data due to the low number of drilled geothermal wells compared to the oil and gas industry [157]. For instance, fewer than 100 geothermal wells were drilled in the United States in 2008 compared to thousands of oil and gas wells [158]. Technology and experience transfer from the oil industry can assist in filling the gap and improving drilling operations in geothermal projects to overcome this knowledge gap and increase drilling efficiency.
Drilling operations in geothermal wells are hampered by the reservoir environment, including lost circulation, well control, and well integrity problems [159]. The choice of drill bit, casing material, drilling mud, and cement formulations are all technically constrained by hard formations and high temperatures. These conditions call for technology innovations to alleviate drilling problems and address high-pressure, high-temperature (HPHT) concerns [160]. High thermal stability drilling fluid formulations are necessary for HPHT settings to endure the high temperatures and avoid problems brought on by fluid degradation. Casing, cement sheaths, and downhole tools can all be harmed by harsh circumstances. Volcanic rocks such as granite, quartzite, granodiorite, and greywacke are typical rock types in geothermal reservoirs. These rocks are recognized for their hardness and abrasiveness, increasing drill bit wear [161]. Increased drill string vibration brought on the formation hardness can result in downhole tool failure [162]. Due to thermally induced stress fatigue, thermal stresses on the casing can result in casing failure when they are greater than the material’s yield stress [163]. Additionally, during drilling and casing operations in geothermal wells, casing failure can happen due to corrosion, wear, and overloading [164]. These damages can impede drilling, well testing, and production activities and, in the worst situations, result in the abandonment of a well.
The examination of crude oil GC-MS is a frequently used technique for identifying different components in crude oil using GC. The flame ionization detector (FID), which produces signals proportional to compound concentrations, is frequently connected to GC. Through the differentiation of components with equal molecular weights but differing polarizabilities, comprehensive two-dimensional gas chromatography (GCxGC) improves resolution. This technique yields more consistent findings than one-dimensional GC because it efficiently detects saturates, aromatics, and chemical interactions. Two styles of a GCxGC plot are shown in Figure 8, which illustrates how well it can identify various components of crude oil, including methyl naphthalenes and their derivatives.
Time-of-flight mass spectrometry helps identify the structure of molecules. While the presence of 25-norhopanes implies extensive biodegradation during petroleum accumulation, the absence of naphthalene in Colombian oil reflects a previous water wash. Additionally, the plot shows that there are two charges of n-alkanes in the reservoir: the first charge biodegrades to produce 25-norhopanes, while the second charge enters the reservoir and undergoes processes of heating and subsidence to produce a lower molecular weight and a minor concentration of hopanes [165].

7. Conclusions and Future Perspectives

This thorough review has critically analysed the difficulties associated with oil and gas well drilling operations and underlined the significance of nanotechnology in the revolution of oil fields. It has become clear from the examination of the issues and suggested solutions that novel strategies are required to increase drilling effectiveness, lower costs, and solve environmental issues. In the fight against these difficulties, nanotechnology has emerged as an area with great potential. Numerous aspects of drilling operations have significantly benefited from the use of nanotechnology in the oil industry. Drilling fluids have advanced features, such as better lubricity using zinc oxide nanoparticles, thermal stability, and filtration capabilities by using silica di-oxide nanoparticles, thanks to the incorporation of these nanoparticles. These developments have improved drilling performance by enhancing wellbore stability, lowering formation damage, and smoothing drilling operations. Moreover, nanotechnology has played a crucial role in the advancement of drilling tools and materials. Nanocomposite materials such as zinc oxide, reinforced with other nanoparticles, have exhibited superior mechanical strength, wear resistance, and thermal stability, thereby enhancing the durability and performance of drilling equipment. The integration of nanosensors and nanorobots has facilitated the real-time monitoring of drilling parameters, empowering operators to make informed decisions, optimize processes, and enhance safety and efficiency.
The integration of nanoparticles to enhance completion fluids, boost oil productivity, and tackle issues such as paraffin removal and enhanced acid treatments is one area where laboratory-scale efforts have demonstrated promise in improving various near-well processes. However, the large-scale production of nanomaterials faces practical challenges, with oil companies’ willingness to participate potentially acting as a limiting factor. There are difficulties in implementing enhanced oil recovery (EOR) operations with nanoparticles, especially when it comes to injectivity problems. Because of the significant volume of nanoparticles employed during EOR, there is a chance that the continuous injection of nanofluids in the water stream would produce injectivity loss in the injector well. Although surfactant-based processes have benefited from the use of nanoparticles, technologies like polymer, alkali, and gas injection have a great deal of room to grow when nanoparticles and nanofluids are added.
The continual injection of nanofluids into the water stream is, in fact, a major problem in the enhanced oil recovery (EOR) process. This raises concerns over potential injectivity loss in the injector well due to numerous types of formation damage. Consequently, it is imperative that the assessment of flow assurance in the laboratory is given top priority to avoid and minimize any formation damage related to nanoparticle injection. While the use of nanoparticles has improved processes based on surfactants, the addition of nanoparticles and nanofluids has the potential to significantly improve technologies such as polymer injection, alkali injection, and gas injection. The handling requirements in surface facilities make it difficult to integrate nanomaterials into polymer, alkali, or gas streams. Robust injection equipment and a complex treatment of generated fluids requiring expensive and time-consuming chemical procedures are required. In order to avoid handling problems that can result in product losses and possible health risks, more research into the process for adding nanoparticles is necessary. Specifically, attention should be paid to investigating methods of incorporating nanoparticles as nanofluids rather than solid systems.
The characterization of reservoirs and improved oil recovery have also benefited from nanotechnology. Improved technologies have been made available to characterize reservoirs through nanoparticle tracers and nanoscale imaging techniques. This has enabled the identification of the best drilling sites and increased productivity. Nanofluids and nanoemulsions have also shown promise in improving oil recovery by increasing fluid mobility, lowering interfacial tension, and altering rock wettability. However, there are obstacles to overcome in applying nanotechnology in the oil industry. Consideration must be given to variables such as stability, the dispersion of nanoparticles, and potential environmental effects. Therefore, future research should develop scalable manufacturing techniques, optimise nanoparticle behaviour within complicated drilling settings, and carefully analyse long-term environmental issues related to nanomaterials. Future possibilities for nanotechnology in the oil industry appear bright. Unlocking the full potential of nanotechnology, further optimizing drilling operations, and addressing current problems will all depend on ongoing research and development activities. Developments will significantly improve drilling efficiency, safety, and environmental sustainability in nanomaterials, nanosensors, and monitoring systems.
Additionally, it is crucial to promote cooperation between regulatory agencies, business, and academia to guarantee the appropriate and long-term integration of nanotechnology in the oil field. A thorough examination of potential hazards and environmental effects, as well as the formulation of rules and standards for the use of nanomaterials, would all be made easier by such collaboration. This thorough review has highlighted the enormous influence of nanotechnology on oil and gas well drilling operations. Nanotechnology has turned out to be a game-changing force in the industry thanks to its capacity to tackle problems and provide creative solutions. Nanotechnology has a wide range of uses, and its use in drilling fluids, innovative materials, real-time monitoring, reservoir characterization, and better oil recovery has shown tremendous promise. Nanotechnology will continue to influence the future of the oil and gas sector by embracing ongoing research, development, and responsible application, enhancing drilling operations, boosting sustainability, and fostering overall industry advancement.
Numerous research and development funds are allocated to the broad field of nanotechnology. Apart from the wide range of uses, such as structural nanomaterials, nanofluids, and nanosensors, nanotechnology also has prospective applications in upstream oil activities. Using nanocatalysts for in situ heavy oil upgrading in reservoirs is one prominent example. Overcoming obstacles can provide significant benefits, such as the high temperatures needed for the catalytic breaking of thick bitumen ends. Still, there are two significant obstacles to the development of these applications. First, some nanomaterials are still somewhat pricey, especially for uses such as enhanced oil recovery (EOR). Second, while introducing nanoparticles into reservoirs, worries regarding the environmental impact surface. Careful life-cycle analyses are necessary, and environmentally friendly injectants such as silica and iron oxide nanoparticles are preferred. Utilizing knowledge from other industries that have thoroughly investigated similar issues is imperative. The upstream oil business appears to have a promising future for nanotechnology if it can adopt notable breakthroughs from various industries.

Author Contributions

Conceptualization, A.M.A. and G.T.C.; writing—original draft preparation, A.M.A.; writing—review and editing, G.T.C.; funding acquisition, A.M.A. and G.T.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ministry of Higher Education, Research and Innovation (MoHERI) of Oman under the Block Funding Program, grant agreement number MoHERI/BFP/ICEM/2023.

Data Availability Statement

Data sharing is not applicable.

Acknowledgments

The authors would like to thank the Ministry of Higher Education, Research and Innovation (MoHERI) Oman for funding this research.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Liu, H.; Jin, X.; Ding, B. Application of nanotechnology in petroleum exploration and development. Pet. Explor. Dev. 2016, 43, 1107–1115. [Google Scholar] [CrossRef]
  2. Abdo, J.; Haneef, M.D. Nanoparticles: Promising solution to overcome stern drilling problems. In Proceedings of the Nanotech Conference and Exhibition, Anaheim, CA, USA, 21–24 June 2010; pp. 6–8. [Google Scholar]
  3. Cheraghian, G.; Khalili Nezhad, S.S.; Bazgir, S. Improvement of thermal stability of polyacryl amide solution used as a nano-fluid in enhanced oil recovery process by nanoclay. Int. J. Nanosci. Nanotechnol. 2015, 11, 201–208. [Google Scholar]
  4. Abdo, J.; Haneef, M.D. Nano-enhanced drilling fluids: Pioneering approach to overcome uncompromising drilling problems. J. Energy Resourse Technol. 2012, 134, 014501. [Google Scholar] [CrossRef]
  5. Devendiran, D.K.; Amirtham, V.A. A review on preparation, characterization, properties and applications of nanofluids. Renew. Sustain. Energy Rev. 2016, 60, 21–40. [Google Scholar] [CrossRef]
  6. Roodbari, P.; Sabbaghi, S. Investigating the properties of modified drilling mud with barite/polyacrylamide nanocomposite. SPE Drill. Complet. 2021, 36, 868–874. [Google Scholar] [CrossRef]
  7. Abdo, J.; Haneef, M.D. Clay nanoparticles modified drilling fluids for drilling of deep hydrocarbon wells. Appl. Clay Sci. 2013, 86, 76–82. [Google Scholar] [CrossRef]
  8. Apaleke, A.S.; Al-Majed, A.; Hossain, M.E. Drilling fluid: State of the art and future trend. In Proceedings of the SPE North Africa Technical Conference and Exhibition, Cairo, Egypt, 20–22 February 2012; p. SPE-149555-MS. [Google Scholar]
  9. Cheraghian, G. Application of nano-particles of clay to improve drilling fluid. Int. J. Nanosci. Nanotechnol. 2017, 13, 177–186. [Google Scholar]
  10. Chekli, L.; Phuntsho, S.; Kim, J.E.; Kim, J.; Choi, J.Y.; Choi, J.-S.; Kim, S.; Kim, J.H.; Hong, S.; Sohn, J.; et al. A comprehensive review of hybrid forward osmosis systems: Performance, applications and future prospects. J. Membr. Sci. 2016, 497, 430–449. [Google Scholar] [CrossRef]
  11. Muhammed, N.S.; Olayiwola, T.; Elkatatny, S.; Haq, B.; Patil, S. Insights into the application of surfactants and nanomaterials as shale inhibitors for water-based drilling fluid: A review. J. Nat. Gas Sci. Eng. 2021, 92, 103987. [Google Scholar] [CrossRef]
  12. Kassim, T.A. Waste Minimization and Molecular Nanotechnology: Toward Total Environmental Sustainability. In Water Pollution: Environmental Impact Assessment of Recycled Wastes on Surface and Ground Waters; Engineering Modeling and Sustainability; Kassim, T.A., Ed.; Springer: Berlin/Heidelberg, Germany, 2005; pp. 191–229. [Google Scholar]
  13. Das, B.; Dutta, D. A review of nanotechnology applications in the oil and gas industries. Resour.-Effic. Technol. 2019, 4, 1–19. [Google Scholar]
  14. Cheraghian, G.; Hendraningrat, L. A review on applications of nanotechnology in the enhanced oil recovery part A: Effects of nanoparticles on interfacial tension. Int. Nano Lett. 2016, 6, 129–138. [Google Scholar] [CrossRef]
  15. Alsedrani, M.Q.; Chala, G.T. Investigation of the Effects of Silica Nanofluid for Enhanced Oil Recovery Applications: CFD Simulation Study. Arab. J. Sci. Eng. 2023, 48, 9139–9158. [Google Scholar] [CrossRef]
  16. Tahr, Z.; Ali, J.A.; Mohammed, A.S. Sustainable aspects behind nano-biodegradable drilling fluids: A critical review. Geoenergy Sci. Eng. 2023, 222, 211443. [Google Scholar] [CrossRef]
  17. Davarpanah, A.; Razmjoo, A.; Mirshekari, B. An overview of management, recycling, and wasting disposal in the drilling operation of oil and gas wells in Iran. Cogent Environ. Sci. 2018, 4, 1537066. [Google Scholar] [CrossRef]
  18. Ko, S.; Huh, C. Use of nanoparticles for oil production applications. J. Pet. Sci. Eng. 2019, 172, 97–114. [Google Scholar] [CrossRef]
  19. Khalil, M.; Jan, B.M.; Tong, C.W.; Berawi, M.A. Advanced nanomaterials in oil and gas industry: Design, application and challenges. Appl. Energy 2017, 191, 287–310. [Google Scholar] [CrossRef]
  20. Cheraghian, G.; Rostami, S.; Afrand, M. Nanotechnology in Enhanced Oil Recovery. Processes 2020, 8, 1073. [Google Scholar] [CrossRef]
  21. Cheraghian, G. Nanoparticles in drilling fluid: A review of the state-of-the-art. J. Mater. Res. Technol. 2021, 13, 737–753. [Google Scholar] [CrossRef]
  22. Davoodi, S.; Al-Shargabi, M.; Wood, D.A.; Rukavishnikov, V.S.; Minaev, K.M. Experimental and field applications of nanotechnology for enhanced oil recovery purposes: A review. Fuel 2022, 324, 124669. [Google Scholar] [CrossRef]
  23. Franco, C.A.; Zabala, R.; Cortés, F.B. Nanotechnology applied to the enhancement of oil and gas productivity and recovery of Colombian fields. J. Pet. Sci. Eng. 2017, 157, 39–55. [Google Scholar] [CrossRef]
  24. Alsaba, M.T.; Al Dushaishi, M.F.; Abbas, A.K. A comprehensive review of nanoparticles applications in the oil and gas industry. J. Pet. Explor. Prod. Technol. 2020, 10, 1389–1399. [Google Scholar] [CrossRef]
  25. Ijaola, A.O.; Farayibi, P.K.; Asmatulu, E. Superhydrophobic coatings for steel pipeline protection in oil and gas industries: A comprehensive review. J. Nat. Gas Sci. Eng. 2020, 83, 103544. [Google Scholar] [CrossRef]
  26. Kianfar, E. Importance & Applications of Nanotechnology; MedDocs Publishers: Reno, NV, USA, 2020. [Google Scholar]
  27. Negin, C.; Ali, S.; Xie, Q. Application of nanotechnology for enhancing oil recovery—A review. Petroleum 2016, 2, 324–333. [Google Scholar] [CrossRef]
  28. Jain, M.; Choudhary, S.; Kumar, V.J.A.E.T. Application of nanotechnology in farm power, machinery and operations: A review. Agric. Eng. Today 2021, 45, 1–9. [Google Scholar] [CrossRef]
  29. Gbadamosi, A.; Junin, R.; Manan, M.; Agi, A.; Oseh, J. Nanotechnology application in chemical enhanced oil recovery: Current opinion and recent advances. In Enhanced Oil Recovery Processes: New Technologies; IntechOpen: London, UK, 2019. [Google Scholar]
  30. Sharma, A.K.; Tiwari, A.K.; Dixit, A.R. Effects of Minimum Quantity Lubrication (MQL) in machining processes using conventional and nanofluid based cutting fluids: A comprehensive review. J. Clean. Prod. 2016, 127, 1–18. [Google Scholar] [CrossRef]
  31. Nagarajan, R. Nanoparticles: Building Blocks for Nanotechnology. In Nanoparticles: Synthesis, Stabilization, Passivation, and Functionalization; American Chemical Society: Washington, DC, USA, 2008; Volume 996, pp. 2–14. [Google Scholar]
  32. Zhe, Z.; An, Y. Nanotechnology for the oil and gas industry—An overview of recent progress. Naotechnol. Rev. 2018, 7, 341–353. [Google Scholar] [CrossRef]
  33. Lau, H.C.; Yu, M.; Nguyen, Q.P. Nanotechnology for oilfield applications: Challenges and impact. J. Pet. Sci. Eng. 2017, 157, 1160–1169. [Google Scholar] [CrossRef]
  34. Saleh, T.A.; Ibrahim, M.A. Advances in functionalized Nanoparticles based drilling inhibitors for oil production. Energy Rep. 2019, 5, 1293–1304. [Google Scholar] [CrossRef]
  35. Salem Ragab, A.M.; Hannora, A.E. A Comparative investigation of nano particle effects for improved oil recovery–experimental work. In Proceedings of the SPE Kuwait Oil and Gas Show and Conference, Mishref, Kuwait, 11–14 October 2015. [Google Scholar]
  36. Sergeev, V.; Tanimoto, K.; Abe, M. Innovative emulsion-suspension systems based on nanoparticles for drilling and well workover operation. In Proceedings of the Abu Dhabi International Petroleum Exhibition and Conference, Abu Dhabi, United Arab Emirates, 11–14 November 2019; p. D031S097R003. [Google Scholar]
  37. Shanmugam, S.; Hari, A.; Pandey, A.; Mathimani, T.; Felix, L.; Pugazhendhi, A. Comprehensive review on the application of inorganic and organic nanoparticles for enhancing biohydrogen production. Fuel 2020, 270, 117453. [Google Scholar] [CrossRef]
  38. Bageri, B.S.; Adebayo, A.R.; Barri, A.; Al Jaberi, J.; Patil, S.; Hussaini, S.R.; Babu, R.S. Evaluation of secondary formation damage caused by the interaction of chelated barite with formation rocks during filter cake removal. J. Pet. Sci. Eng. 2019, 183, 106395. [Google Scholar] [CrossRef]
  39. Mohamed, A.; Benaafi, M.; Elkatatny, S.; Bageri, B. Effect of high-density water-based drilling fluid on the mechanical properties of the drilled formation in horizontal wells. In Proceedings of the ARMA US Rock Mechanics/Geomechanics Symposium, New York, NY, USA, 23–26 June 2019; p. ARMA-2019-1683. [Google Scholar]
  40. Bageri, B.S.; Adebayo, A.R.; Al Jaberi, J.; Patil, S. Effect of perlite particles on the filtration properties of high-density barite weighted water-based drilling fluid. Powder Technol. 2020, 360, 1157–1166. [Google Scholar] [CrossRef]
  41. Wang, B.; Sun, J.; Shen, F.; Li, W.; Zhang, W. Mechanism of wellbore instability in continental shale gas horizontal sections and its water-based drilling fluid countermeasures. Nat. Gas Ind. B 2020, 7, 680–688. [Google Scholar] [CrossRef]
  42. You, L.; Kang, Y.; Chen, Z.; Chen, Q.; Yang, B. Wellbore instability in shale gas wells drilled by oil-based fluids. Int. J. Rock Mech. Min. Sci. 2014, 72, 294–299. [Google Scholar] [CrossRef]
  43. Bloys, B.; Davis, N.; Smolen, B.; Bailey, L.; Houwen, O.; Reid, P.; Sherwood, J.; Fraser, L.; Hodder, M.; Montrouge, F.J.O.R. Designing and managing drilling fluid. Oilfield Rev. 1994, 6, 33–43. [Google Scholar]
  44. Pašić, B.; Gaurina-Međimurec, N.; Mijić, P.; Medved, I. Experimental Research of Shale Pellet Swelling in Nano-Based Drilling Muds. Energies 2020, 13, 6246. [Google Scholar] [CrossRef]
  45. Mukhametshin, V.S.; Zeigman, Y.V.; Andreev, A.V. Rapid assessment of deposit production capacity for determination of nanotechnologies application efficiency and necessity to stimulate their development. Nanotekhnologii V Stroit. 2017, 9, 20–34. [Google Scholar] [CrossRef]
  46. Nagarajan, K.; Ramanujam, N.; Sanjay, M.; Siengchin, S.; Surya Rajan, B.; Sathick Basha, K.; Madhu, P.; Raghav, G. A comprehensive review on cellulose nanocrystals and cellulose nanofibers: Pretreatment, preparation, and characterization. Polym. Compos. 2021, 42, 1588–1630. [Google Scholar] [CrossRef]
  47. Khan, F.; Kesarwani, H.; Kataria, G.; Mittal, G.; Sharma, S. Application of titanium dioxide nanoparticles for the design of oil well cement slurry—A study based on compressive strength, setting time and rheology. J. Adhes. Sci. Technol. 2023, 37, 1666–1682. [Google Scholar] [CrossRef]
  48. Afolabi, R.O.; Yusuf, E.O. Nanotechnology and global energy demand: Challenges and prospects for a paradigm shift in the oil and gas industry. J. Pet. Explor. Prod. Technol. 2019, 9, 1423–1441. [Google Scholar] [CrossRef]
  49. Afolabi, R.O.; Orodu, O.D.; Efeovbokhan, V.E. Properties and application of Nigerian bentonite clay deposits for drilling mud formulation: Recent advances and future prospects. Appl. Clay Sci. 2017, 143, 39–49. [Google Scholar] [CrossRef]
  50. Aftab, A.; Ismail, A.R.; Ibupoto, Z.H.; Akeiber, H.; Malghani, M.G.K. Nanoparticles based drilling muds a solution to drill elevated temperature wells: A review. Renew. Sustain. Energy Rev. 2017, 76, 1301–1313. [Google Scholar] [CrossRef]
  51. Agista, M.N.; Guo, K.; Yu, Z. A State-of-the-Art Review of Nanoparticles Application in Petroleum with a Focus on Enhanced Oil Recovery. Appl. Sci. 2018, 8, 871. [Google Scholar] [CrossRef]
  52. Ali, M.; Jarni, H.H.; Aftab, A.; Ismail, A.R.; Saady, N.M.C.; Sahito, M.F.; Keshavarz, A.; Iglauer, S.; Sarmadivaleh, M. Nanomaterial-Based Drilling Fluids for Exploitation of Unconventional Reservoirs: A Review. Energies 2020, 13, 3417. [Google Scholar] [CrossRef]
  53. Li, X.-L.; Jiang, G.-C.; Xu, Y.; Deng, Z.-Q.; Wang, K. A new environmentally friendly water-based drilling fluids with laponite nanoparticles and polysaccharide/polypeptide derivatives. Pet. Sci. 2022, 19, 2959–2968. [Google Scholar] [CrossRef]
  54. Ali, J.A.; Kolo, K.; Manshad, A.K.; Mohammadi, A.H. Recent advances in application of nanotechnology in chemical enhanced oil recovery: Effects of nanoparticles on wettability alteration, interfacial tension reduction, and flooding. Egypt. J. Pet. 2018, 27, 1371–1383. [Google Scholar] [CrossRef]
  55. Ali, J.A.; Kalhury, A.M.; Sabir, A.N.; Ahmed, R.N.; Ali, N.H.; Abdullah, A.D. A state-of-the-art review of the application of nanotechnology in the oil and gas industry with a focus on drilling engineering. J. Pet. Sci. Eng. 2020, 191, 107118. [Google Scholar] [CrossRef]
  56. El-Masry, J.F.; Bou-Hamdan, K.F.; Abbas, A.H.; Martyushev, D.A. A Comprehensive Review on Utilizing Nanomaterials in Enhanced Oil Recovery Applications. Energies 2023, 16, 691. [Google Scholar] [CrossRef]
  57. Bennetzen, M.V.; Mogensen, K. Novel applications of nanoparticles for future enhanced oil recovery. In Proceedings of the International Petroleum Technology Conference, Kuala Lumpur, Malaysia, 10 December 2014; p. IPTC-17857-MS. [Google Scholar]
  58. Seyedmohammadi, J. The effects of drilling fluids and environment protection from pollutants using some models. Model. Earth Syst. Environ. 2017, 3, 23. [Google Scholar] [CrossRef]
  59. Razali, S.Z.; Yunus, R.; Abdul Rashid, S.; Lim, H.N.; Mohamed Jan, B. Review of biodegradable synthetic-based drilling fluid: Progression, performance and future prospect. Renew. Sustain. Energy Rev. 2018, 90, 171–186. [Google Scholar] [CrossRef]
  60. Bera, A.; Belhaj, H. Application of nanotechnology by means of nanoparticles and nanodispersions in oil recovery—A comprehensive review. J. Nat. Gas Sci. Eng. 2016, 34, 1284–1309. [Google Scholar] [CrossRef]
  61. Borisov, A.S.; Husein, M.; Hareland, G. A field application of nanoparticle-based invert emulsion drilling fluids. J. Nanopart. Res. 2015, 17, 340. [Google Scholar] [CrossRef]
  62. Cheraghian, G.; Wu, Q.; Mostofi, M.; Li, M.-C.; Afrand, M.; Sangwai, J.S. Effect of a novel clay/silica nanocomposite on water-based drilling fluids: Improvements in rheological and filtration properties. Colloids Surf. A Physicochem. Eng. Asp. 2018, 555, 339–350. [Google Scholar] [CrossRef]
  63. Zamora-Ledezma, C.; Narváez-Muñoz, C.; Guerrero, V.H.; Medina, E.; Meseguer-Olmo, L. Nanofluid Formulations Based on Two-Dimensional Nanoparticles, Their Performance, and Potential Application as Water-Based Drilling Fluids. ACS Omega 2022, 7, 20457–20476. [Google Scholar] [CrossRef] [PubMed]
  64. Kök, M.V.; Bal, B. Effects of silica nanoparticles on the performance of water-based drilling fluids. J. Pet. Sci. Eng. 2019, 180, 605–614. [Google Scholar] [CrossRef]
  65. Smith, S.R.; Rafati, R.; Sharifi Haddad, A.; Cooper, A.; Hamidi, H. Application of aluminium oxide nanoparticles to enhance rheological and filtration properties of water based muds at HPHT conditions. Colloids Surf. A Physicochem. Eng. Asp. 2018, 537, 361–371. [Google Scholar] [CrossRef]
  66. Singh, R.; Singh, D.; Pandey, G. Comprehensive review on nanoparticles and its applications in petroleum industry. AIP Conf. Proc. 2023, 2521, 030011. [Google Scholar]
  67. Pervaiz, S.; Kannan, S.; Kishawy, H.A. An extensive review of the water consumption and cutting fluid based sustainability concerns in the metal cutting sector. J. Clean. Prod. 2018, 197, 134–153. [Google Scholar] [CrossRef]
  68. Vryzas, Z.; Kelessidis, V.C. Nano-Based Drilling Fluids: A Review. Energies 2017, 10, 540. [Google Scholar] [CrossRef]
  69. Sabet, M.; Hosseini, S.; Zamani, A.; Hosseini, Z.; Soleimani, H. Application of nanotechnology for enhanced oil recovery: A review. In Defect and Diffusion Forum; Trans Tech Publications: Wallerau, Switzerland, 2016; pp. 149–156. [Google Scholar]
  70. Yang, J.; Ji, S.; Li, R.; Qin, W.; Lu, Y. Advances of Nanotechnologies in Oil and Gas Industries. Energy Explor. Exploit. 2015, 33, 639–657. [Google Scholar] [CrossRef]
  71. Al-Shargabi, M.; Davoodi, S.; Wood, D.A.; Al-Musai, A.; Rukavishnikov, V.S.; Minaev, K.M. Nanoparticle applications as beneficial oil and gas drilling fluid additives: A review. J. Mol. Liq. 2022, 352, 118725. [Google Scholar] [CrossRef]
  72. Peng, B.; Tang, J.; Luo, J.; Wang, P.; Ding, B.; Tam, K.C. Applications of nanotechnology in oil and gas industry: Progress and perspective. Can. J. Chem. Eng. 2018, 96, 91–100. [Google Scholar] [CrossRef]
  73. Lee, J.; Babadagli, T. Comprehensive review on heavy-oil emulsions: Colloid science and practical applications. Chem. Eng. Sci. 2020, 228, 115962. [Google Scholar] [CrossRef]
  74. Hemath, M.; Mavinkere Rangappa, S.; Kushvaha, V.; Dhakal, H.N.; Siengchin, S. A comprehensive review on mechanical, electromagnetic radiation shielding, and thermal conductivity of fibers/inorganic fillers reinforced hybrid polymer composites. Polym. Compos. 2020, 41, 3940–3965. [Google Scholar] [CrossRef]
  75. Muhammed, N.S.; Olayiwola, T.; Elkatatny, S. A review on clay chemistry, characterization and shale inhibitors for water-based drilling fluids. J. Pet. Sci. Eng. 2021, 206, 109043. [Google Scholar] [CrossRef]
  76. Hussain Qaiser, M.S.; Ahmad, I.; Ahmad, S.R.; Afzal, M.; Qayyum, A. Assessing Heavy Metal Contamination in Oil and Gas Well Drilling Waste and Soil in Pakistan. Pol. J. Environ. Stud. 2019, 28, 785–793. [Google Scholar] [CrossRef]
  77. Alkalbani, A.M.; Chala, G.T.; Zar Myint, M.T. Insightful study on the effect of zinc oxide nanoparticle diameter on the rheology of water base mud at elevated temperature. J. Pet. Sci. Eng. 2022, 217, 110878. [Google Scholar] [CrossRef]
  78. Alkalbani, A.M.; Chala, G.T.; Myint, M.T.Z. Experimental investigation of rheological properties of water-base mud with zinc oxide nanoparticles using response surface methodology. J. Pet. Sci. Eng. 2022, 208, 109781. [Google Scholar] [CrossRef]
  79. Alkalbani, A.K.; Chala, G.T.; Alkalbani, A.M. Experimental investigation of the rheological properties of water base mud with silica nanoparticles for deep well application. Ain Shams Eng. J. 2023, 14, 102147. [Google Scholar] [CrossRef]
  80. Singh, D.; Ruhil, D.; Khandelwal, H.; Rawat, H.; Aggarwal, H.; Ranjan, A.; Thakur, N.K. Investigation of rheological and filtration properties of water-based drilling mud using commercially available additives. Mater. Today Proc. 2022, 68, 1003–1010. [Google Scholar] [CrossRef]
  81. Hajiabadi, S.H.; Aghaei, H.; Kalateh-Aghamohammadi, M.; Shorgasthi, M. An overview on the significance of carbon-based nanomaterials in upstream oil and gas industry. J. Pet. Sci. Eng. 2020, 186, 106783. [Google Scholar] [CrossRef]
  82. Gautam, R.; Sahai, M.; Kumar, S. Recent advances in application of nanomaterials as additives for drilling fluids. Energy Sources Part A Recovery Util. Environ. Eff. 2020, 1–24. [Google Scholar] [CrossRef]
  83. Hussein, A.K. Applications of nanotechnology in renewable energies—A comprehensive overview and understanding. Renew. Sustain. Energy Rev. 2015, 42, 460–476. [Google Scholar] [CrossRef]
  84. Hoelscher, K.P.; De Stefano, G.; Riley, M.; Young, S. Application of nanotechnology in drilling fluids. In Proceedings of the SPE International Oilfield Nanotechnology Conference and Exhibition, Noordwijk, The Netherlands, 12 June 2012; p. SPE-157031-MS. [Google Scholar]
  85. Koh, J.K.; Lai, C.W.; Johan, M.R.; Gan, S.S.; Chua, W.W. Recent advances of modified polyacrylamide in drilling technology. J. Pet. Sci. Eng. 2022, 215, 110566. [Google Scholar] [CrossRef]
  86. Lysakova, E.; Skorobogatova, A.; Neverov, A.; Pryazhnikov, M.; Zhigarev, V.; Voronin, A.; Shabanova, K.; Minakov, A. Comparative analysis of the effect of single-walled and multi-walled carbon nanotube additives on the properties of hydrocarbon-based drilling fluids. Colloids Surf. A Physicochem. Eng. Asp. 2023, 678, 132434. [Google Scholar] [CrossRef]
  87. Mangadlao, J.D.; Cao, P.; Advincula, R.C. Smart cements and cement additives for oil and gas operations. J. Pet. Sci. Eng. 2015, 129, 63–76. [Google Scholar] [CrossRef]
  88. Shafiei, M.; Kazemzadeh, Y.; Martyushev, D.A.; Dai, Z.; Riazi, M. Effect of chemicals on the phase and viscosity behavior of water in oil emulsions. Sci. Rep. 2023, 13, 4100. [Google Scholar] [CrossRef]
  89. Agwu, O.E.; Akpabio, J.U.; Ekpenyong, M.E.; Inyang, U.G.; Asuquo, D.E.; Eyoh, I.J.; Adeoye, O.S. A comprehensive review of laboratory, field and modelling studies on drilling mud rheology in high temperature high pressure (HTHP) conditions. J. Nat. Gas Sci. Eng. 2021, 94, 104046. [Google Scholar] [CrossRef]
  90. Abdullah, A.H.; Ridha, S.; Mohshim, D.F.; Yusuf, M.; Kamyab, H.; Krishna, S.; Maoinser, M.A. A comprehensive review of nanoparticles: Effect on water-based drilling fluids and wellbore stability. Chemosphere 2022, 308, 136274. [Google Scholar] [CrossRef]
  91. Jiang, G.; Dong, T.; Cui, K.; He, Y.; Quan, X.; Yang, L.; Fu, Y. Research status and development directions of intelligent drilling fluid technologies. Pet. Explor. Dev. 2022, 49, 660–670. [Google Scholar] [CrossRef]
  92. Nemati, M.M.; Saboori, R.; Sabbaghi, S. Hydrogen sulfide removal using various metal oxide nanocomposite from drilling fluid: Optimization, kinetic and adsorption isotherms. Beilstein Arch. 2019, 2019, 133. [Google Scholar]
  93. Vryzas, Z.; Mahmoud, O.; Nasr-El-Din, H.A.; Kelessidis, V.C. Development and testing of novel drilling fluids using Fe2O3 and SiO2 nanoparticles for enhanced drilling operations. In Proceedings of the International Petroleum Technology Conference, Doha, Qatar, 6–9 December 2015. [Google Scholar]
  94. Mahmoud, O.; Nasr-El-Din, H.A.; Vryzas, Z.; Kelessidis, V.C. Nanoparticle-based drilling fluids for minimizing formation damage in HP/HT applications. In Proceedings of the SPE International Conference and Exhibition on Formation Damage Control, Lafayette, LA, USA, 24–26 February 2016. [Google Scholar]
  95. Okoro, E.E.; Oladejo, B.R.; Sanni, S.E.; Obomanu, T.; Ibe, A.A.; Orodu, O.D.; Olawole, O.C. A new model for curbing filtrate loss in dynamic application of nano-treated aqueous mud systems. Adv. Nano Res. 2020, 9, 59–67. [Google Scholar]
  96. Oseh, J.O.; Mohd, N.M.N.A.; Gbadamosi, A.O.; Agi, A.; Blkoor, S.O.; Ismail, I.; Igwilo, K.C.; Igbafe, A.I. Polymer nanocomposites application in drilling fluids: A review. Geoenergy Sci. Eng. 2023, 222, 211416. [Google Scholar] [CrossRef]
  97. Sirajunnisa, A.R.; Surendhiran, D. Algae—A quintessential and positive resource of bioethanol production: A comprehensive review. Renew. Sustain. Energy Rev. 2016, 66, 248–267. [Google Scholar] [CrossRef]
  98. Witherspoon, T.A. Nano-Steel-Co. Available online: https://www.nanosteelco.com/ (accessed on 10 October 2023).
  99. Skandon, G.; Singhal, A. Nanomaterials; Recent Advances in Technology and Industry; Taylor & Francis: Boca Raton, FL, USA, 2004. [Google Scholar]
  100. Arizona, U.o. SRC Engineering Research Center for Environmentally Benign Semiconductor Manufacturing (ERC). Available online: http://erc.arizona.edu/ (accessed on 23 August 2023).
  101. Sun, Y.; Yang, D.; Shi, L.; Wu, H.; Cao, Y.; He, Y.; Xie, T. Properties of Nanofluids and Their Applications in Enhanced Oil Recovery: A Comprehensive Review. Energy Fuels 2020, 34, 1202–1218. [Google Scholar] [CrossRef]
  102. Thibaud, M.; Chi, H.; Zhou, W.; Piramuthu, S. Internet of Things (IoT) in high-risk Environment, Health and Safety (EHS) industries: A comprehensive review. Decis. Support Syst. 2018, 108, 79–95. [Google Scholar] [CrossRef]
  103. Umaraw, P.; Verma, A.K. Comprehensive review on application of edible film on meat and meat products: An eco-friendly approach. Crit. Rev. Food Sci. Nutr. 2017, 57, 1270–1279. [Google Scholar] [CrossRef]
  104. Yekeen, N.; Padmanabhan, E.; Idris, A.K.; Chauhan, P.S. Nanoparticles applications for hydraulic fracturing of unconventional reservoirs: A comprehensive review of recent advances and prospects. J. Pet. Sci. Eng. 2019, 178, 41–73. [Google Scholar] [CrossRef]
  105. Zhang, W.; Zheng, Q.; Ashour, A.; Han, B. Self-healing cement concrete composites for resilient infrastructures: A review. Compos. Part B Eng. 2020, 189, 107892. [Google Scholar] [CrossRef]
  106. Vipulanandan, C.; Krishnamoorti, R.; Mohammed, A.; Boncan, V.; Narvaez, G.; Head, B.; Pappas, J. Iron nanoparticle modified smart cement for real time monitoring of ultra deepwater oil well cementing applications. In Proceedings of the Offshore Technology Conference, Houston, TX, USA, 7 May 2015; p. OTC-25842-MS. [Google Scholar]
  107. Thakkar, A.; Raval, A.; Chandra, S.; Shah, M.; Sircar, A. A comprehensive review of the application of nano-silica in oil well cementing. Petroleum 2020, 6, 123–129. [Google Scholar] [CrossRef]
  108. Zhu, D.; Peng, S.; Zhao, S.; Wei, M.; Bai, B. Comprehensive Review of Sealant Materials for Leakage Remediation Technology in Geological CO2 Capture and Storage Process. Energy Fuels 2021, 35, 4711–4742. [Google Scholar] [CrossRef]
  109. Shah, S.S.; Shaikh, M.N.; Khan, M.Y.; Alfasane, M.A.; Rahman, M.M.; Aziz, M.A. Present status and future prospects of jute in nanotechnology: A review. Chem. Rec. 2021, 21, 1631–1665. [Google Scholar] [CrossRef]
  110. Ejileugha, C.; Ezejiofor, A.N.; Ezealisiji, K.M.; Orisakwe, O.E. Metal oxide nanoparticles in oil drilling: Aquatic toxicological concerns. J. Hazard. Mater. Adv. 2022, 7, 100116. [Google Scholar] [CrossRef]
  111. Fakoya, M.F.; Shah, S.N. Emergence of nanotechnology in the oil and gas industry: Emphasis on the application of silica nanoparticles. Petroleum 2017, 3, 391–405. [Google Scholar] [CrossRef]
  112. Fakoya, M.F.; Shah, S.N. Effect of silica nanoparticles on the rheological properties and filtration performance of surfactant-based and polymeric fracturing fluids and their blends. SPE Drill. Complet. 2018, 33, 100–114. [Google Scholar] [CrossRef]
  113. Ikram, R.; Jan, B.M.; Ahmad, W.; Sidek, A.; Khan, M.; Kenanakis, G. Rheological Investigation of Welding Waste-Derived Graphene Oxide in Water-Based Drilling Fluids. Materials 2022, 15, 8266. [Google Scholar] [CrossRef]
  114. Jia, X.; Zhao, X.; Chen, B.; Egwu, S.B.; Huang, Z. Polyanionic cellulose/hydrophilic monomer copolymer grafted silica nanocomposites as HTHP drilling fluid-loss control agent for water-based drilling fluids. Appl. Surf. Sci. 2022, 578, 152089. [Google Scholar] [CrossRef]
  115. Jiang, G.; Sun, J.; He, Y.; Cui, K.; Dong, T.; Yang, L.; Yang, X.; Wang, X. Novel Water-Based Drilling and Completion Fluid Technology to Improve Wellbore Quality During Drilling and Protect Unconventional Reservoirs. Engineering 2022, 18, 129–142. [Google Scholar] [CrossRef]
  116. Khorshidia, S.; Sabbaghia, S.; Saboorib, R. Removal of hydrogen sulfide from water-based drilling fluid using titania/carbon nanotubes nano-hybrid: Optimization, kinetics, and isotherms. Desalin. Water Treat. 2020, 200, 154–166. [Google Scholar] [CrossRef]
  117. Misbah, B.; Sedaghat, A.; Rashidi, M.; Sabati, M.; Vaidyan, K.; Ali, N.; Omar, M.A.A.; Hosseini Dehshiri, S.S. Friction reduction of Al2O3, SiO2, and TiO2 nanoparticles added to non-Newtonian water based mud in a rotating medium. J. Pet. Sci. Eng. 2022, 217, 110927. [Google Scholar] [CrossRef]
  118. Moghadam, I.P.; Afrand, M.; Hamad, S.M.; Barzinjy, A.A.; Talebizadehsardari, P. Curve-fitting on experimental data for predicting the thermal-conductivity of a new generated hybrid nanofluid of graphene oxide-titanium oxide/water. Phys. A Stat. Mech. Appl. 2020, 548, 122140. [Google Scholar] [CrossRef]
  119. Mohamed, A.; Salehi, S.; Ahmed, R. Significance and complications of drilling fluid rheology in geothermal drilling: A review. Geothermics 2021, 93, 102066. [Google Scholar] [CrossRef]
  120. Technologies, I. Breaking Boundaries. Driving Innovation. Available online: https://unctad.org/system/files/official-document/tir2023_en.pdf (accessed on 2 April 2023).
  121. Razali, S.Z.; Yunus, R.; Kania, D.; Rashid, S.A.; Ngee, L.H.; Abdulkareem-Alsultan, G.; Jan, B.M. Effects of morphology and graphitization of carbon nanomaterials on the rheology, emulsion stability, and filtration control ability of drilling fluids. J. Mater. Res. Technol. 2022, 21, 2891–2905. [Google Scholar] [CrossRef]
  122. Ng, L.Y.; Mohammad, A.W.; Leo, C.P.; Hilal, N. Polymeric membranes incorporated with metal/metal oxide nanoparticles: A comprehensive review. Desalination 2013, 308, 15–33. [Google Scholar] [CrossRef]
  123. Srivastava, V.; Beg, M.; Sharma, S.; Choubey, A.K. Application of manganese oxide nanoparticles synthesized via green route for improved performance of water-based drilling fluids. Appl. Nanosci. 2021, 11, 2247–2260. [Google Scholar] [CrossRef]
  124. Franco, C.A.; Franco, C.A.; Zabala, R.D.; Bahamón, Í.; Forero, Á.; Cortés, F.B. Field Applications of Nanotechnology in the Oil and Gas Industry: Recent Advances and Perspectives. Energy Fuels 2021, 35, 19266–19287. [Google Scholar] [CrossRef]
  125. Baig, U.; Gondal, M.A.; Dastageer, M.A. Oil-water separation using surface engineered superhydrophobic and superoleophilic membrane for the production of clean water. J. Water Process Eng. 2022, 45, 102473. [Google Scholar] [CrossRef]
  126. Bardhan, A.; Khan, F.; Kesarwani, H.; Vats, S.; Sharma, S.; Kumar, S. Performance Evaluation of Novel Silane Coated Nanoparticles as an Additive for High-Performance Drilling Fluid Applications. In Proceedings of the International Petroleum Technology Conference, Bangkok, Thailand, 28 February–2 March 2023. [Google Scholar]
  127. Boopathi, S.; Davim, J.P. Applications of Nanoparticles in Various Manufacturing Processes. In Sustainable Utilization of Nanoparticles and Nanofluids in Engineering Applications; IGI Global: Hershey, PA, USA, 2023; pp. 1–31. [Google Scholar]
  128. Chala, G.T.; Sulaiman, S.A.; Japper-Jaafar, A. Flow start-up and transportation of waxy crude oil in pipelines-A review. J. Non-Newton. Fluid Mech. 2018, 251, 69–87. [Google Scholar] [CrossRef]
  129. Hussain, A.; Emadi, H.; Botchway, K. How nanoparticles have ameliorated the challenges in drilling operations. J. Pet. Sci. Eng. 2021, 197, 107931. [Google Scholar] [CrossRef]
  130. Ikram, R.; Mohamed Jan, B.; Vejpravova, J.; Choudhary, M.I.; Zaman Chowdhury, Z. Recent Advances of Graphene-Derived Nanocomposites in Water-Based Drilling Fluids. Nanomaterials 2020, 10, 2004. [Google Scholar] [CrossRef]
  131. Mohanty, U.S.; Aftab, A.; Awan, F.U.R.; Ali, M.; Yekeen, N.; Keshavarz, A.; Iglauer, S. Toward Improvement of Water-Based Drilling Mud via Zirconia Nanoparticle/API Bentonite Material. Energy Fuels 2022, 36, 12116–12125. [Google Scholar] [CrossRef]
  132. Okyere, M.S. Corrosion Protection for the Oil and Gas Industry: Pipelines, Subsea Equipment, and Structures; CRC Press: Boca Raton, FL, USA, 2019. [Google Scholar]
  133. Khan, F.S.A.; Mubarak, N.M.; Tan, Y.H.; Khalid, M.; Karri, R.R.; Walvekar, R.; Abdullah, E.C.; Nizamuddin, S.; Mazari, S.A. A comprehensive review on magnetic carbon nanotubes and carbon nanotube-based buckypaper for removal of heavy metals and dyes. J. Hazard. Mater. 2021, 413, 125375. [Google Scholar] [CrossRef]
  134. Medhi, S.; Chowdhury, S.; Gupta, D.K.; Mazumdar, A. An investigation on the effects of silica and copper oxide nanoparticles on rheological and fluid loss property of drilling fluids. J. Pet. Explor. Prod. Technol. 2020, 10, 91–101. [Google Scholar] [CrossRef]
  135. Sircar, A.; Rayavarapu, K.; Bist, N.; Yadav, K.; Singh, S. Applications of nanoparticles in enhanced oil recovery. Pet. Res. 2022, 7, 77–90. [Google Scholar] [CrossRef]
  136. Kaur, R.; Sharma, S.K.; Tripathy, S.K. Chapter 7—Advantages and Limitations of Environmental Nanosensors. In Advances in Nanosensors for Biological and Environmental Analysis; Deep, A., Kumar, S., Eds.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 119–132. [Google Scholar]
  137. Joshi, A.; Appold, M.S.; Nunn, J.A. Evaluation of solitary waves as a mechanism for oil transport in poroelastic media: A case study of the South Eugene Island field, Gulf of Mexico basin. Mar. Pet. Geol. 2012, 37, 53–69. [Google Scholar] [CrossRef]
  138. Murtaza, M.; Alarifi, S.A.; Kamal, M.S.; Onaizi, S.A.; Al-Ajmi, M.; Mahmoud, M. Experimental Investigation of the Rheological Behavior of an Oil-Based Drilling Fluid with Rheology Modifier and Oil Wetter Additives. Molecules 2021, 26, 4877. [Google Scholar] [CrossRef] [PubMed]
  139. Al-Saba, M.; Amadi, K.; Al-Hadramy, K.; Dushaishi, M.; Al-Hameedi, A.; Alkinani, H. Experimental investigation of bio-degradable environmental friendly drilling fluid additives generated from waste. In Proceedings of the SPE International Conference and Exhibition on Health, Safety, Security, Environment, and Social Responsibility, Abu Dhabi, United Arab Emirates, 16–18 April 2018. [Google Scholar]
  140. Ahmadi, M.H.; Ghazvini, M.; Alhuyi Nazari, M.; Ahmadi, M.A.; Pourfayaz, F.; Lorenzini, G.; Ming, T.J.I.J.o.E.R. Renewable energy harvesting with the application of nanotechnology: A review. Int. J. Energy Res. 2019, 43, 1387–1410. [Google Scholar] [CrossRef]
  141. Ghaderi, S.; Clarkson, C.; Kaviani, D. Investigation of Primary Recovery in Low-Permeability Oil Formations: A Look at the Cardium Formation, Alberta (Canada). J. Oil Gas Sci. Technol.–Rev. D’ifp Energ. Nouv. 2014, 69, 1155–1170. [Google Scholar] [CrossRef]
  142. Al-Hameedi, A.T.T.; Alkinani, H.H.; Alkhamis, M.M.; Dunn-Norman, S. Utilizing a new eco-friendly drilling mud additive generated from wastes to minimize the use of the conventional chemical additives. J. Pet. Explor. Prod. Technol. 2020, 10, 3467–3481. [Google Scholar] [CrossRef]
  143. Shekdar, A.V. Sustainable solid waste management: An integrated approach for Asian countries. Waste Manag. 2009, 29, 1438–1448. [Google Scholar] [CrossRef] [PubMed]
  144. Taran, M.; Safaei, M.; Karimi, N.; Almasi, A. Benefits and application of nanotechnology in environmental science: An overview. J Biointerface Res. Appl. Chem. 2021, 11, 7860–7870. [Google Scholar]
  145. Sharmin, T.; Khan, N.R.; Akram, M.S.; Ehsan, M.M. A State-of-the-Art Review on Geothermal Energy Extraction, Utilization, and Improvement Strategies: Conventional, Hybridized, and Enhanced Geothermal Systems. Int. J. Thermofluids 2023, 18, 100323. [Google Scholar] [CrossRef]
  146. Kiran, R.; Dansena, P.; Salehi, S.; Rajak, V.K. Application of machine learning and well log attributes in geothermal drilling. Geothermics 2022, 101, 102355. [Google Scholar] [CrossRef]
  147. Reinsch, T.; Regenspurg, S.; Feldbusch, E.; Saadat, A.; Huenges, E.; Erbas, K.; Zimmermann, G.; Henninges, J. Operations. Reverse cleanout in a geothermal well: Analysis of a failed coiled-tubing operation. SPE Prod. 2015, 30, 312–320. [Google Scholar] [CrossRef]
  148. Basfar, S.; Elkatatny, S. Micronized calcium carbonate to enhance water-based drilling fluid properties. Sci. Rep. 2023, 13, 18295. [Google Scholar] [CrossRef]
  149. Al-Ghanimi, G.F.; Al-Zubaidi, N.S. Effect of Ferric Oxide and Magnesium Oxide Nanoparticles on Iraqi Bentonite Performance in Water Based Drilling Fluid. Assoc. Arab Univ. J. Eng. Sci. 2020, 27, 14–23. [Google Scholar] [CrossRef]
  150. Gasser, M.; Mahmoud, O.; Ibrahim, F.; Abadir, M. Using artificial intelligence techniques in modeling and predicting the rheological behavior of nano-based drilling fluids. In Proceedings of the International Conference on Offshore Mechanics and Arctic Engineering, Virtual, 21–30 June 2021; American Society of Mechanical Engineers: New York, NY, USA, 2021; p. V010T011A070. [Google Scholar]
  151. Gupta, D.; Chauhan, V.; Kumar, R. A comprehensive review on synthesis and applications of molybdenum disulfide (MoS2) material: Past and recent developments. Inorg. Chem. Commun. 2020, 121, 108200. [Google Scholar] [CrossRef]
  152. Alihosseini, A.; Hassan Zadeh, A.; Monajjemi, M.; Nazary Sarem, M. Effectiveness of Alumina Nanoparticles in Improving the Rheological Properties of Water-Based Drilling Mud. Iran. J. Oil Gas Sci. Technol. 2021, 10, 12–27. [Google Scholar]
  153. Mehmood, A.; Khan, F.S.A.; Mubarak, N.M.; Mazari, S.A.; Jatoi, A.S.; Khalid, M.; Tan, Y.H.; Karri, R.R.; Walvekar, R.; Abdullah, E.C.; et al. Carbon and polymer-based magnetic nanocomposites for oil-spill remediation—A comprehensive review. Environ. Sci. Pollut. Res. 2021, 28, 54477–54496. [Google Scholar] [CrossRef]
  154. Minakov, A.V.; Mikhienkova, E.I.; Voronenkova, Y.O.; Neverov, A.L.; Zeer, G.M.; Zharkov, S.M. Systematic experimental investigation of filtration losses of drilling fluids containing silicon oxide nanoparticles. J. Nat. Gas Sci. Eng. 2019, 71, 102984. [Google Scholar] [CrossRef]
  155. Rafati, R.; Smith, S.R.; Sharifi Haddad, A.; Novara, R.; Hamidi, H. Effect of nanoparticles on the modifications of drilling fluids properties: A review of recent advances. J. Pet. Sci. Eng. 2018, 161, 61–76. [Google Scholar] [CrossRef]
  156. Ramasamy, J.; Amanullah, M. Nanocellulose for oil and gas field drilling and cementing applications. J. Pet. Sci. Eng. 2020, 184, 106292. [Google Scholar] [CrossRef]
  157. Rana, A.; Khan, I.; Saleh, T.A. Advances in Carbon Nanostructures and Nanocellulose as Additives for Efficient Drilling Fluids: Trends and Future Perspective—A Review. Energy Fuels 2021, 35, 7319–7339. [Google Scholar] [CrossRef]
  158. Saleh, T.A. Nanotechnology in Oil and Gas Industries; Springer: Cham, Switzerland, 2018. [Google Scholar]
  159. Al-Yasiri, M.S.; Al-Sallami, W.T. How the drilling fluids can be made more efficient by using nanomaterials. Am. J. Nano Res. Appl. 2015, 3, 41–45. [Google Scholar]
  160. Boul, P.J.; Ajayan, P.M. Nanotechnology research and development in upstream oil and gas. Energy Technol. 2020, 8, 1901216. [Google Scholar] [CrossRef]
  161. Ciambelli, P.; La Guardia, G.; Vitale, L. Chapter 7—Nanotechnology for green materials and processes. In Studies in Surface Science and Catalysis; Basile, A., Centi, G., Falco, M.D., Iaquaniello, G., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; Volume 179, pp. 97–116. [Google Scholar]
  162. Debnath, B.K.; Saha, U.K.; Sahoo, N. A comprehensive review on the application of emulsions as an alternative fuel for diesel engines. Renew. Sustain. Energy Rev. 2015, 42, 196–211. [Google Scholar] [CrossRef]
  163. Karakosta, K.; Mitropoulos, A.C.; Kyzas, G.Z. A review in nanopolymers for drilling fluids applications. J. Mol. Struct. 2021, 1227, 129702. [Google Scholar] [CrossRef]
  164. Medhi, S.; Chowdhury, S.; Sangwai, J.S.; Gupta, D.K. Effect of Al2O3 nanoparticle on viscoelastic and filtration properties of a salt-polymer-based drilling fluid. Energy Sources Part A Recovery Util. Environ. Eff. 2023, 45, 2385–2397. [Google Scholar] [CrossRef]
  165. Jia, B.; Xian, C.; Tsau, J.-S.; Zuo, X.; Jia, W. Status and Outlook of Oil Field Chemistry-Assisted Analysis during the Energy Transition Period. Energy Fuels 2022, 36, 12917–12945. [Google Scholar] [CrossRef]
Figure 1. Number of publications on the application of nanotechnology in drilling in the past two decades [16].
Figure 1. Number of publications on the application of nanotechnology in drilling in the past two decades [16].
Energies 17 00798 g001
Figure 2. Percentage of the conducted investigations into nanoparticles across the oil and gas industry [24].
Figure 2. Percentage of the conducted investigations into nanoparticles across the oil and gas industry [24].
Energies 17 00798 g002
Figure 3. Wellbore instability mechanisms [44].
Figure 3. Wellbore instability mechanisms [44].
Energies 17 00798 g003
Figure 4. (a) Shale Membrane Tester (SMT) test apparatus set-up. (b) The pressure in shale changes when silica nanoparticles are injected using a water-based drilling fluid [84].
Figure 4. (a) Shale Membrane Tester (SMT) test apparatus set-up. (b) The pressure in shale changes when silica nanoparticles are injected using a water-based drilling fluid [84].
Energies 17 00798 g004
Figure 5. Schematic of oil droplet removal process using surface-engineered SPM-NPs [125].
Figure 5. Schematic of oil droplet removal process using surface-engineered SPM-NPs [125].
Energies 17 00798 g005
Figure 6. (a) Magnetic nanoparticle-based heating (left) and magnetic field oscillation generator (right). (b) Partially melted wax pieces flowing with crude oil in the pipe, as the SPM-NP-imbedded “nano-paint” is heated up using magnetic field oscillation [132].
Figure 6. (a) Magnetic nanoparticle-based heating (left) and magnetic field oscillation generator (right). (b) Partially melted wax pieces flowing with crude oil in the pipe, as the SPM-NP-imbedded “nano-paint” is heated up using magnetic field oscillation [132].
Energies 17 00798 g006
Figure 7. Graphical representation of okra powder as an eco-friendly additive in water-based drilling fluids. Increase in plastic viscosity, yield point, apparent viscosity, gel strength (a,b), and decrease in fluid loss and mud cake thickness (c).
Figure 7. Graphical representation of okra powder as an eco-friendly additive in water-based drilling fluids. Increase in plastic viscosity, yield point, apparent viscosity, gel strength (a,b), and decrease in fluid loss and mud cake thickness (c).
Energies 17 00798 g007
Figure 8. GCxGC illustrations of (a) Eagle Ford oil and (b) Colombian oil [165].
Figure 8. GCxGC illustrations of (a) Eagle Ford oil and (b) Colombian oil [165].
Energies 17 00798 g008
Table 1. Percentage of NPs in oil and gas applications [24].
Table 1. Percentage of NPs in oil and gas applications [24].
MaterialPercentage
Aluminium Oxide9%
Copper Oxide2%
Graphene Based5%
Iron Oxide5%
Magnesium Oxide5%
MWCNT3%
Silicon Dioxide23%
Titanium Oxide4%
Nickle Oxide2%
Zinc Oxide3%
Others39%
Table 2. Improved property by means of NPs.
Table 2. Improved property by means of NPs.
Property ImprovementPercentage
Improving Foam Stability 3%
Increase Oil Recovery29%
Improving Rheological Properties 12%
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Alkalbani, A.M.; Chala, G.T. A Comprehensive Review of Nanotechnology Applications in Oil and Gas Well Drilling Operations. Energies 2024, 17, 798. https://doi.org/10.3390/en17040798

AMA Style

Alkalbani AM, Chala GT. A Comprehensive Review of Nanotechnology Applications in Oil and Gas Well Drilling Operations. Energies. 2024; 17(4):798. https://doi.org/10.3390/en17040798

Chicago/Turabian Style

Alkalbani, Alhaitham M., and Girma T. Chala. 2024. "A Comprehensive Review of Nanotechnology Applications in Oil and Gas Well Drilling Operations" Energies 17, no. 4: 798. https://doi.org/10.3390/en17040798

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop