8.7 C
New York
Saturday, April 20, 2024
Home Blog

The Importance of Rigorous Safety Protocols on Construction Sites

Construction sites can be really dangerous and that is why governments have set certain safety protocols that will protect the workers as much as possible. Implementing and enforcing rigorous safety protocols is not merely a regulatory requirement; it is an absolute necessity for safeguarding the well-being of personnel and ensuring project success.

Construction projects involve a multitude of activities that expose workers to various dangers. These hazards include:

  • Falls: Falls from heights, scaffolding, ladders, and roofs remain a leading cause of construction site fatalities.
  • Falling Objects: Tools, materials, and debris pose a constant threat if not secured properly.
  • Electrocution: Improper electrical wiring, exposed conductors, and contact with energized equipment can lead to serious electrical injuries.
  • Trench Collapses: Improper trench shoring and excavation practices can result in catastrophic cave-ins.
  • Heavy Equipment: Operating heavy machinery like cranes, forklifts, and bulldozers requires vigilance to prevent collisions and crushing accidents.
  • Exposure to Hazardous Materials: Construction materials like asbestos, lead, and silica dust can pose severe health risks if not handled appropriately.
  • Noise Hazards: Continuous exposure to loud noise from machinery can lead to hearing loss.

There are many more reasons why these protocols are in place and why they need to be followed fully.

The workers are in danger

The main reason why we have so many protocols on construction sites is because they will keep you safe to some extent, and many of them have lowered fatalities and other injuries that happen to workers. You may find some of those protocols to be boring, and we understand that, but they need to be done regularly so that we can limit injuries and other accidents to a minimum.

Imagine that you skip a safety protocol that you should follow, and then an accident with you or some coworkers happens. The guilt that you will feel cannot be taken away since you cannot go into the past. Maybe some of your coworkers will blame you for what happened, even though that may not be the case. Imagine all the trauma you will have to endure and you could have avoided all of this if you just did what was asked of you.

Adhering to the law

Even though safety precautions should be the main reason why you follow these protocols, most businesses and people do so because they have to. The government has set hefty fines for those who do not adhere to safety protocols. For some companies, the only way they will do what they are told is by forcing them to.

Even though it would be great that they feel the need to do these things themselves because they protect their people and the firm, they still hesitate either because it costs them more money or because they do not care. Also, if the company fails to comply with the set safety rules, it can lose its reputation which will hinder its growth and income. 

Legal Help

Nearly 150,000 construction workers sustain injuries every year, as reported by the Bureau of Labor Statistics. Injuries often result from falls, but they can also happen as a result of physical contact with machinery. By following the rules, wearing safety equipment, and never putting yourself in harm’s way, you can ensure that your workplace is safe.

But the reality is that no matter how careful you are, you run the risk of suffering a major accident on a construction site. That is where a legal professional comes into play, according to a Staten Island Personal Injury Lawyer. Construction attorneys serve large construction businesses, employees, property owners, sureties, and many more customers.

Indeed, a construction attorney can turn out to be necessary for everyone dealing with construction at some point in their lives. They can help protect the victim’s rights and fight for the compensation they deserve that can cover both the physical and emotional pain they have endured. 

The company is saving money

Many people think that they are wasting time doing all of these protocols on the worksite and that it takes time from actual work, which subsequently leads to a waste of money. This is not the right approach to looking at this issue. You need to realize that you will lose much more money if you do not comply with regulations.

If the government catches you, you will have to pay a lot of fines and could also go to court. If one of your employees gets hurt, they will get much more money out of your company than they would if you had done everything by the book. So it can turn out to be much more expensive for you if you decide to skip the safety protocols. 

Better reputation

A bad reputation is something that every company must look to avoid. There are so many ways you can destroy your credibility, which will lead to your company struggling and maybe even completely going under. One of those many ways is that it gets out there that your construction sites are not safe and that they offer dangerous working conditions. If you decide to skip the safety protocol, you could seriously jeopardize your company’s image, especially if injuries or fatalities were involved. By taking this issue seriously, you can avoid future headaches and continue to grow. 

Enhanced productivity

What some owners think is that if they follow all of these protocols, efficiency will take a hit but we completely disagree with that statement. When the workers on a construction site ensure that everything is up to code, they will not have to worry about the many dangers that are out there on the worksite. When they feel more relaxed, they can go about their business and concentrate on it. They can rely on the equipment they have because it has been properly tested just hours before.

Also, because there are fewer injuries on the job site, there will be fewer stops. Imagine someone getting seriously hurt on the site. Everyone immediately seizes all the work they were doing to see what has happened, and that is normal because they worry about their colleagues and they want to help.

Higher morale

When you regularly listen to the guidance of the government and do all the relevant things they demand, you will manage to create a worksite that is as safe as it can be, which means that fewer accidents and injuries will happen. Because accidents happen rarely, or at least less often than on other construction sites, you will create a great atmosphere among them.

It is always great to be able to lift the spirits of your employees and by nurturing a safe space for them, you can expect great things. Nobody wants to feel in danger all the time so you should look for ways to make your workers feel safe and one way is through safety protocols. 

There is a moral and practical as well as a legal need to ensure the safety of everyone on building projects. Aside from saving lives, prioritizing safety boosts productivity and creates a great work atmosphere.

Conclusion

Prioritizing safety on construction sites is not merely a regulatory requirement; it is a moral and economic imperative. Rigorous safety protocols safeguard the well-being of workers, promote project success, and enhance a company’s reputation. By fostering a safety-centric culture and continuously striving for improvement, construction companies can build a safer future for their workforce and contribute to a more sustainable and responsible construction industry.

Civil Engineers and Building Information Modelling (BIM)

Building Information Modelling (BIM) constitutes a structured methodology for the generation and administration of data about a building, infrastructure asset, or facility throughout its entire life cycle. The major outcome of this process is the Building Information Model (BIM), which serves as a digital representation of all the important features of the constructed entity.

This will usually involve the digital model of the architectural features, structural elements, electrical and mechanical services, etc. Notably, this model is progressively developed and refined through the collaborative incorporation and updating of information during the project’s execution. The prominent features of BIM in the construction industry are highlighted as follows;

  1. BIM as a Digital Repository: Building Information Modelling (BIM) leverages a digital model to serve as a comprehensive repository for data and information about a building, infrastructure, or facility throughout its entire lifecycle. This model necessitates continuous access, enrichment, and modification to maintain its accuracy and reflect project evolution.
  2. BIM as an Information Management Process: Beyond the digital model itself, BIM encompasses a structured process, or rather, a network of interconnected activities. This process focuses on managing the information embedded within these models to maximize their utility and optimize project outcomes.
  3. BIM-Driven Collaboration: A cornerstone principle of BIM is the emphasis on collaborative workflows. To ensure the information models remain current and usable, all stakeholders must engage in collaborative efforts at designated stages of the process, adhering to established protocols for data contribution and exchange.

Building Information Modeling (BIM) has revolutionized the construction industry, transforming how civil engineers approach design, construction, and project management. This technology fosters collaboration, optimizes workflows, and enhances decision-making throughout a project’s lifecycle. This article discusses the effects of BIM on civil engineering practices.

Building information modelling BIM

BIM vs. Traditional CAD

Traditionally, civil engineers relied on 2D drawings for design and communication. These drawings, while informative, often lacked the necessary depth and detail for comprehensive project visualization. BIM introduces a paradigm shift by creating intelligent 3D models involving all aspects of a civil engineering project. These models integrate architectural, structural, and MEP (Mechanical, Electrical, Plumbing) components, providing a holistic view of the infrastructure being designed.

Therefore, Building Information Modeling (BIM) fundamentally diverges from the conventional Computer-Aided Design (CAD) approach. This distinction lies in the concept of a singular source of truth. In BIM, all modifications are conducted directly on the central BIM model itself. As a consequence, individual plan derivations such as sections, views, and floor plans no longer necessitate independent updates when alterations are introduced.

The core distinction also manifests in the nature of the data being manipulated. BIM models are comprised of intelligent objects, imbued with semantic structure. These objects, representing elements like walls or slabs, transcend mere geometric representation. Conversely, CAD drawings solely depict geometry through interconnected lines and arcs, lacking the inherent intelligence of BIM objects.

The advantages of BIM in the construction industry extend well beyond the foundational concept of a centralized model. Well-established benefits, such as automated clash detection and code checking, are now widely recognized and employed within the industry. Additionally, visualization techniques leveraging Augmented Reality (AR) and Virtual Reality (VR) technologies are gaining traction, offering enhanced project understanding and communication.

Perhaps the most significant impact of BIM lies in its ability to foster exemplary collaboration amongst stakeholders involved in the design, construction, and operational phases of a project. As BIM workflows become increasingly integrated into standard practice, these well-known benefits are continuously being exploited and optimized, leading to significant improvements in project delivery.

BIM and Civil Engineers

Listed below are the benefits of BIM for civil engineers.

  • Enhanced Collaboration: BIM facilitates seamless collaboration between civil engineers, architects, contractors, and other stakeholders. The 3D model serves as a central repository of information, enabling real-time communication and clash detection. This collaborative environment minimizes errors and omissions, leading to a more efficient design process.
  • Improved Design Accuracy: BIM allows for the creation of highly detailed and accurate models, incorporating precise information about elements like roadways, bridges, utilities, and drainage systems. This approach reduces the likelihood of errors during construction, minimizing rework and associated costs.
  • Optimized Project Management: BIM empowers civil engineers to leverage the data embedded within the 3D model for effective project management. The model allows for quantity takeoffs, cost estimation, and scheduling optimization. This data-driven approach fosters informed decision-making throughout the construction process.
  • Enhanced Sustainability: BIM facilitates the evaluation of a project’s environmental impact during the design phase. The model allows for simulations to assess factors like material usage, energy consumption, and lifecycle costs. This enables civil engineers to design sustainable infrastructure that minimizes environmental footprint and optimizes resource utilization.
  • Reduced Risk and Improved Safety: BIM facilitates clash detection, a process where potential conflicts between different design elements are identified within the 3D model. This proactive approach allows for early resolution of these conflicts, minimizing risks associated with rework and construction delays. Additionally, BIM can be used to simulate construction sequencing, promoting improved worker safety on-site.

BIM Applications in Civil Engineering

  • Transportation Infrastructure: BIM plays a crucial role in the design and construction of roads, bridges, and tunnels. 3D models enable detailed analysis of traffic flow, structural integrity, and constructability.
  • Water Resources Management: BIM facilitates the design of water treatment plants, dams, and irrigation systems. The 3D model allows for accurate modelling of water flow, hydraulic simulations, and construction sequencing.
  • Site Development: BIM is instrumental in planning and designing site layouts, grading plans, and utility networks. The 3D model facilitates visualization of site constraints, underground infrastructure, and potential constructability challenges.

Challenges and the Road Ahead

Despite its immense benefits, BIM adoption in civil engineering faces challenges. These include the initial investment in software and training, the need for standardized data exchange formats, and the integration of BIM workflows with existing design practices. However, the construction industry is rapidly recognizing the transformative power of BIM. As technology advances, software becomes more user-friendly, and standardized data formats emerge, BIM is poised for even greater integration within civil engineering workflows.

Conclusion

BIM is no longer a novelty but a necessity for civil engineers in today’s competitive construction landscape. By embracing BIM, civil engineers can create efficient, sustainable, and cost-effective infrastructure projects while fostering a collaborative and data-driven approach to construction. As BIM continues to evolve, civil engineers will be at the forefront of shaping the future of the built environment.

Corrosion of Steel Structures

Structural steel is a major material in modern construction that offers exceptional strength, versatility, and cost-effectiveness. However, its susceptibility to corrosion can present a significant challenge to structural engineers and asset managers. Corrosion in steel structures is not only an aesthetic issue, but can impact the structures’ stability, longevity, and safety. The deterioration of steel, often termed corrosion, can be primarily understood as an electrochemical process that occurs sequentially in the presence of water and oxygen.

The initial phase involves the targeted degradation of specific surface regions, known as anodes. At these anodic sites, ferrous ions dissolve and are released into the surrounding environment. Simultaneously, electrons depart from the anode and travel through the metallic structure to adjoining cathodic locations on the surface. At these cathodic sites, the electrons interact with oxygen and water, forming hydroxyl ions.

Subsequently, these hydroxyl ions react with the ferrous ions originating from the anode to produce ferrous hydroxide. Finally, this ferrous hydroxide undergoes further oxidation in the presence of air, ultimately transforming into hydrated ferric oxide, commonly recognized as red rust.

The following equation describes the sum of these reactions:
4Fe + 3O2 + 2H2O = 2Fe2O3H2O
(iron/steel) + (oxygen) + (water) = rust

Two critical observations can therefore be drawn regarding steel corrosion:

  1. The deterioration of iron or steel, known as corrosion, necessitates the concurrent presence of both water and oxygen. In the absence of either element, corrosion ceases to occur.
  2. The entirety of the corrosion process transpires at the anode; the cathode experiences no corrosion.

However, with time, polarization effects come into play. These effects, such as the accumulation of corrosion products on the surface, hinder the continuation of the corrosion process. Subsequently, new and reactive anodic sites may form, permitting further corrosion to take place. Over extended periods, the loss of metal becomes relatively uniform across the surface, leading to a phenomenon typically described as general corrosion.

Different Forms of Steel Corrosion

Steel corrosion manifests itself in various forms, each posing unique challenges:

  • Uniform Corrosion: This most common form involves the gradual, near-uniform deterioration of the entire exposed steel surface. While seemingly less damaging initially, it can significantly reduce the load-bearing capacity of a structure over time.
  • Pitting Corrosion: This localized attack creates deep, often invisible pits that compromise structural integrity disproportionately to the overall material loss. It is particularly problematic due to its rapid progression and difficulty in detection.
  • Galvanic Corrosion: This occurs when two dissimilar metals, like steel and copper, are in electrical contact in a corrosive environment. The more “noble” metal (copper) acts as the cathode, accelerating the corrosion of the less noble metal (steel).
  • Crevice Corrosion: This localized attack occurs in confined spaces between the steel surface and another material, such as a gasket or sealant. The stagnant electrolyte within this crevice promotes a highly corrosive environment.
  • Stress Corrosion Cracking (SCC): This form combines the effects of tensile stress and a corrosive environment, leading to the rapid propagation of cracks within the steel.

Rate of Steel Corrosion

The primary factors influencing the rate of steel corrosion in atmospheric environments can be categorized as follows:

(1) Wetness Duration: This metric refers to the proportional time during which the steel surface remains wet due to factors such as precipitation or condensation. Consequently, unprotected steel in arid environments, exemplified by heated buildings, experiences minimal corrosion owing to the limited availability of water.

(2) Atmospheric Contaminants: The type and concentration of airborne pollutants and contaminants, including sulfur dioxide, chlorides, and dust particles, significantly impact corrosion rates.

(3) Sulfates: These originate from sulfur dioxide gas, a byproduct of fossil fuel combustion (e.g., sulfur-laden coal and oil). Sulfur dioxide reacts with atmospheric moisture to form sulfuric and sulfurous acids. Industrial environments are particularly susceptible to high sulfur dioxide concentrations.

(4) Chlorides: Primarily found in marine environments, chlorides reach peak concentrations in coastal regions, exhibiting a rapid decline further inland. However, the presence of de-icing salts used on roadways can introduce chlorides into non-coastal environments.

It is noteworthy that within a specific geographic location, corrosion rates can exhibit significant variability due to the sheltering effects of structures and prevailing wind patterns. Therefore, for practical purposes, the immediate “microclimate” surrounding the steel structure dictates its corrosion rate.

This section explores the corrosion rates of steel in a range of United Kingdom environments, expressed in micrometres per year (μm/year). Note: 1 μm (micrometre) is equivalent to 0.001 mm (millimetre).

  • Rural Atmospheric: Unobstructed, inland environments with minimal pollution generally exhibit low steel corrosion rates, typically below 50 μm/year.
  • Industrial Atmospheric: Inland environments characterized by air pollution experience moderate corrosion rates, typically ranging from 40 to 80 μm/year, with variations influenced by sulfur dioxide (SO2) levels.
  • Marine Atmospheric: Within the UK, a broad 2-kilometre coastal strip is considered a marine environment. Steel corrosion rates in this zone typically fall between 50 and 100 μm/year, heavily influenced by proximity to the seawater.
  • Marine Industrial Atmospheric: Coastal environments with significant pollution exposure exhibit the highest corrosion rates, ranging from 50 to 150 μm/year.
  • Seawater Immersion: Four distinct vertical zones are typically encountered in tidal waters, each with varying corrosion rates:
    • Splash Zone (Above High Tide): This zone experiences the most severe corrosion, with an average rate of approximately 75 μm/year.
    • Tidal Zone (High Tide to Low Tide): Often covered by marine organisms, this zone exhibits lower corrosion rates, averaging around 35 μm/year.
    • Low-Water Zone (Just Below Low Tide): Corrosion rates in this narrow band are comparable to the splash zone.
    • Permanent Immersion Zone (Low-Water Level to Seabed): This zone exhibits the lowest corrosion rates, averaging around 35 μm/year.
  • Freshwater Immersion: Steel corrosion rates in freshwater environments are generally lower than those observed in saltwater, typically ranging from 30 to 50 μm/year.

Impact of Steel Corrosion

The effects of steel corrosion are far-reaching and pose significant economic and safety concerns in structural steel works:

  • Structural Failure: Severe corrosion can compromise the structural integrity of steel beams, columns, and connections, potentially leading to catastrophic failure.
  • Aesthetics and Property Value: Visible corrosion not only detracts from the visual appeal of a structure but can also decrease its market value.
  • Maintenance Costs: The continuous battle against corrosion necessitates ongoing inspection, repair, and replacement of affected steel components, leading to substantial financial burdens.

Corrosion Protection of Steel Structures

For most common applications, specifying cost-effective protective treatments for structural steelwork becomes a straightforward process when the factors influencing durability are understood. The primary consideration lies in recognizing and defining the corrosivity of the environment where the structure will reside. This understanding is very important for selecting an appropriate protective system.

Many structures are situated in relatively low-risk environments, necessitating minimal protective treatment. Conversely, exposure to an aggressive environment necessitates a more durable protective system, potentially requiring maintenance to ensure extended service life. Striking the optimal balance involves combining proper surface preparation with suitable coating materials to achieve the desired durability at the lowest possible cost.

Modern practices, applied in accordance with relevant industry standards, offer the opportunity to attain the specific protection requirements for various structures. Numerous standards exist to aid in drafting protection specifications. One of the most significant is ISO 12944, titled “Paints and Varnishes – Corrosion Protection of Steel Structures by Protective Paint Systems.” This comprehensive standard, published in eight parts, serves as a valuable resource when crafting protection specifications for structural steelwork.

Strategies for Mitigating Corrosion

A multi-pronged approach is required in mitigating steel corrosion and ensuring the long-term serviceability of steel structures:

  • Material Selection: Selecting steel alloys with enhanced corrosion resistance, such as weathering steels or those with higher chromium content, can be very beneficial.
  • Protective Coatings: Applying paints, galvanizing (zinc coating), or using cathodic protection systems (electrical current application to suppress corrosion) act as barriers against the corrosive environment.
  • Design Considerations: Design features that minimize water accumulation, such as proper drainage, ventilation, and avoiding crevices, can significantly retard corrosion progression.
  • Regular Inspection and Maintenance: Proactive inspection programs that identify and address corrosion early are essential to prevent catastrophic failures.

Metallic Coating

Four primary methods are employed for applying metallic coatings to steel surfaces, each offering distinct advantages:

  • Hot-Dip Galvanizing: This process involves immersing the steel in molten zinc, resulting in the formation of a zinc-iron alloy layer that provides excellent corrosion resistance.
  • Thermal (Metal) Spraying: This technique utilizes a high-temperature source to melt a metallic wire, which is then sprayed onto the prepared steel surface. Thermal spraying offers the flexibility to apply a variety of coating materials but may exhibit lower consistency in coating thickness compared to other methods.
  • Electroplating: While not typically used for structural steelwork due to limitations in achievable coating thickness, electroplating is a suitable method for coating fittings, fasteners, and other smaller steel components. This process involves depositing a thin layer of metal onto the steel surface using an electrical current.
  • Sherardizing: Similar to electroplating, sherardizing is not commonly employed for structural steelwork. This process involves diffusing a zinc powder coating onto the steel surface at elevated temperatures. It finds application in protecting smaller steel items.

It is important to note that the effectiveness of metallic coatings in protecting against corrosion is primarily influenced by the type of coating metal chosen and its applied thickness. The specific method of application, with the exception of thermal metal spraying due to potential thickness variations, has a lesser impact on overall corrosion resistance. In this article, we will focus on hot-dip galvanising as a method of metallic coating.

Hot-dip Galvanizing

Hot-dip galvanizing is the most prevalent method for applying a metallic coating to structural steel. This process adheres to the specifications outlined in ISO 1461, which mandates a minimum zinc coating weight of 610 g/m² (equivalent to a minimum average thickness of 85 micrometres) for sections no less than 6 millimetres thick.

Hot-dip galvanised structural sections for corrosion protection
Hot-dip galvanised structural sections for corrosion protection

The process of hot-dip galvanizing is as follows:

  1. Degreasing: Any surface contaminants like oil or grease are removed using appropriate degreasing agents.
  2. Surface Preparation: All rust and scale are eliminated from the steel through acid pickling. In some cases, this stage may be preceded by blast-cleaning to enhance surface roughness and remove scale. However, blast-cleaned surfaces invariably undergo subsequent pickling with inhibited hydrochloric acid.
  3. Flux Application: Following cleaning, the steel is immersed in a fluxing agent. This step ensures optimal contact between the molten zinc and the steel during the dipping process.
  4. Galvanizing: The cleaned and fluxed steel is submerged in a molten zinc bath maintained at approximately 450°C. At this elevated temperature, the steel undergoes a reactive process with the molten zinc, resulting in the formation of a series of zinc-iron alloy layers on its surface.
  5. Zinc Layer Formation: Upon extraction from the bath, a layer of virtually pure zinc deposits on top of the pre-formed alloy layers.
Hot dip galvanising

As the zinc solidifies, it adopts a characteristic crystalline metallic sheen, commonly referred to as spangling. The final thickness of the galvanized coating is influenced by several factors, including:

  • Size and thickness of the steel workpiece
  • Surface preparation technique employed (abrasive blast cleaning typically yields thicker coatings)
  • Steel composition

The inherent nature of hot-dip galvanizing, being a dipping process, imposes limitations on the size of components that can be treated. For workpieces exceeding the dimensions of the galvanizing bath, a technique called double dipping can be employed. This involves sequentially dipping one end of the item before the other.

In many applications, hot-dip galvanized steel offers sufficient protection without further treatment. However, for enhanced durability, particularly in specific atmospheric environments, or for aesthetic purposes, paint coatings can be applied over the galvanized surface.

Thermal (metal) Spraying

Thermal (metal) spraying presents an alternative approach for applying a metallic coating to structural steelwork. This method offers the flexibility of utilizing either zinc or aluminium as the coating material, typically supplied in powder or wire form.

The core process involves feeding the metal through a specialized spray gun equipped with a heat source, which can be either an oxy-fuel flame or an electric arc. The heat source melts the metal, transforming it into molten globules that are then propelled onto the previously blast-cleaned steel surface using compressed air.

It is important to note that, unlike hot-dip galvanizing, no alloying occurs between the coating and the steel substrate. The resulting coating consists of overlapping metallic platelets with a porous structure. This necessitates subsequent sealing of the pores

The adhesion of thermally sprayed metal coatings to steel is primarily attributed to mechanical interlocking. Consequently, achieving an optimal bond necessitates applying the coating to a thoroughly cleaned and roughened steel surface. Blast-cleaning with a coarse abrasive grit is the standard practice for surface preparation.

Typical coating thicknesses for thermally sprayed aluminium range from 150 to 200 micrometres, while zinc coatings typically fall between 100 and 150 micrometres.

A significant advantage of thermal metal spraying lies in its versatility. Unlike hot-dip galvanizing, which has limitations on workpiece size due to the dipping process, thermal spraying can be performed either in workshops or directly on-site, accommodating components of any size. Additionally, as the steel surface remains cool during the spraying process, concerns regarding thermal distortion are mitigated.

Design guidance for objects intended for thermal spraying can be found in BS 4479-7. However, it is important to acknowledge that thermal metal spraying is a considerably more expensive option compared to hot-dip galvanizing.

Paint Coatings

Painting remains the primary method for safeguarding structural steelwork from corrosion. Paints consist of a carefully balanced mixture of three key components: pigments, binders, and solvents. Application to steel surfaces can be achieved through various methods; however, all methods result in an initial wet film. As the solvent evaporates, a film-forming process occurs, leaving behind a dry film composed of pigments and binders adhering to the surface.

Common classification systems for paints categorize them based on either pigmentation or binder type. Primers designed for steel are usually classified according to the predominant corrosion-inhibiting pigments incorporated into their formulation. Examples include zinc phosphate and metallic zinc.

These inhibitive pigments can be formulated with various binder resins, resulting in variations like zinc phosphate alkyd primers or zinc phosphate epoxy primers. Intermediate and finishing coats are typically classified based on their binders (e.g., epoxy build coats, vinyl finishes, urethane finishes) or their pigments.

Paint application typically involves layering multiple coats, each serving a specific purpose.

The primer, applied directly to the cleaned steel surface, serves the dual function of wetting the surface and ensuring optimal adhesion for subsequent layers. Primers formulated for steel surfaces often possess additional corrosion-inhibiting properties.

Intermediate coats, also known as undercoats, contribute to building the overall film thickness of the paint system. This may necessitate the application of several coats.

Finishing coats provide the frontline defence against environmental elements while also determining the final aesthetic appearance in terms of gloss, colour, and other visual characteristics.

Compatibility between the various superimposed coats within a painting system is crucial for optimal performance. Additionally, vulnerable areas of the structure require the application of extra coats, known as stripe coats, to achieve the minimum required thickness. As a general best practice, it is recommended to source all paints within a system from the same manufacturer.

The chosen application method and prevailing environmental conditions significantly impact the quality and longevity of the applied coating. While other methods like dipping may be employed, the standard application methods for paint on structural steelwork include brush, roller, conventional air-spray, and airless spray techniques.

Conclusion

Steel remains a vital material in modern construction. However, the ever-present threat of corrosion necessitates a thorough understanding of the underlying mechanisms, the various forms of attack, and the potential consequences. By implementing a comprehensive corrosion mitigation strategy, engineers and asset managers can ensure the safety, longevity, and economic viability of steel structures.

Construction and Safety: 6 Legal Tips to Follow When Working On-Site

Construction sites are always bustling with activity as hundreds of employees work round the clock to give architectural dreams a shape and form. But behind the din of machinery and buzz of construction, there is a very stark reality: construction reigns as one of the most dangerous industries for workplace­ accidents, injuries, and deaths.

The Occupational Safety and He­alth Administration (OSHA) shows construction accounts for a significant portion of occupational fatalities yearly, with falls, ele­ctrocutions, and struck by objects being among the leading cause­s. Thus, ensuring construction worker safety isn’t just a moral imperative­, but a legal requireme­nt avoiding costly legal issues, protecting human live­s, and the reputation of construction companies.

construction site

Here are the six (6) legal tips to follow when working on-site.

  1. Obtain the Necessary Permits and Licenses

A critical step is getting the necessary permits and licenses before starting any construction project. This involves getting the right paperwork with legal authorizations from relevant authorities, as well as confirmation about the project’s compliance with local regulation standards.

However, failure to acquire permits has really bad legal implications, such as hefty fines, project shutdown, and serious legal action. The required types of permits will range from zoning regulations and building codes to environmental protection protocols, depending on the nature and scope of the project in question.

  1. Comply with OSHA Regulations

OSHA ensures employees’ work environments are safe and healthful through the definition of rules and directions that are in line with many different industries, including construction sites. These policies aim at lessening dangers, stopping harm, and ensuring the well-being of the workers. Ideally, the OSHA regulations provide the basis for:

Legal Obligations and Worker Protection

Complying with OSHA requirements is not just about meeting legal obligations but also about giving the construction worker top consideration for his safety and health. Complying ensures that employers clearly demonstrate strong commitments to providing safe working environments.

Importance of Proper Safety Protocols

Having proper safety protocols under OSHA compliance means spotting the prospective hazards at a construction site, evaluating potential risks, and finding proper ways to control and mitigate those risks. It should cover broader areas for safety protocols, including operating equipment, handling hazardous material, and emergency response procedures.

Key Components of OSHA Compliance

Safety and health, according to OSHA rules, need to be approached with much detail. For example, constant checking helps in the determination of potential risks. Besides, all the workers should be provided with training and a complete set of means of safety needed to carry out the hazards. Moreover, the employer must make records of training, inspections, and incidents to display compliance with OSHA regulations.

  1. Maintain Proper Documentation

Maintaining prope­r documentation proves vital for construction firms to show compliance with safe­ty rules, conduct thorough inspections, and tackle any incide­nts or disputes occurring on-site properly. Furthermore­, precise paperwork is a vital instrume­nt defending corporate me­asures during court cases; thus, it is pivotal in lesse­ning legal dangers. To ensure compliance and mitigate­ these risks, employe­rs should establish clear procedure­s for documenting various safety measures, and any other relevant documentation pertaining to safety protocols and regulatory requirements.

In the unfortunate event of an injury due to negligence or unsafe working conditions, seeking legal counsel from experienced personal injury lawyers is essential. According to the legal team at Prochaska, Howell & Prochaska, experienced lawyers can help injured workers navigate the legal complexities and pursue rightful compensation. They will also ensure fair settlements on behalf of the injured party through negotiation and gathering essential evidence to build a strong case. 

  1. Implement Proper Safety Training Programs

Comprehensive training prepares workers with all the knowledge that they are supposed to have, in addition to the skills necessary to undertake jobs in ways that are safe and minimize risks of accidents and injuries. Thus, the lack of proper training endangers the workers and may place construction companies at risk of legal liability. Employers are to design and develop sound safety training programs, which must be custom-built and in line with the prevailing hazards on site.

For employe­rs to develop good safety training programs, ide­ntifying the unique risks workers may face­ is vital. Conducting thorough assessments and understanding construction re­gulations is essential. Furthermore, the content should cover those risks like fall protection, e­quipment use, and eme­rgency plans.

Implementation Strategies

Planning safety training programs ne­eds thoughtful preparation and exe­cution. Firms should schedule freque­nt training sessions to ensure all worke­rs get proper instruction. Skilled traine­rs ought to be hired to delive­r training, providing hands-on demonstrations and applied guidance. 

Ongoing Evaluation and Improvement

Safety training programs require ongoing appraisal and e­nhancement. This involves ge­tting worker feedback to gauge­ training effectivene­ss and pinpoint improvement areas. Training programs should be­ revised as require­d to tackle any shortcomings or emerging risks. Furthermore, through commitme­nt to continual learning and progress, employe­rs can ensure their safe­ty training programs stay effective and re­levant over time.

Importance of Comprehensive Safety Training

Comprehensive safe­ty training is vital for several reasons. For instance, it:

  • E­nsures workers are compe­tent and prepared to handle the challenges and hazards that come with the job.
  • It helps in safeguarding against injuries, cre­ating a safer workplace for e­veryone.
  • It helps avoid legal implications associated with inadequate training.
  1. Ensure Compliance with Building Codes and Regulations

Construction projects must follow building rule­s. These regulations se­t standards for materials, designs, and construction methods. In addition, the­y protect people inside­ and outside the buildings. However, not following the rule­s can lead to fines, delays, or stopping work. So, construction companie­s need to know the late­st codes for their projects. The­y must plan and build with these rules from the­ start, as it prevents legal issue­s and promotes safety during construction.

  1. Establish Clear Contracts and Agreements

Putting clear contracts, agreements, and de­als in writing is key for building projects. These­ papers show who does what, when, and how. Furthermore, the­y map out the plan, due dates, pay te­rms, and safety rules. Clear pape­rs also help prevent mix-ups by giving guidance and settling disputes. Hence, by stating e­veryone’s jobs and rights, builders avoid le­gal issues, facilitating work to move on smoothly. Plus, ope­n deals build trust among team membe­rs, contributing to the overall success of the projects.

Kee­ping sites safe matters for workers, passersby, and the environment – le­gally and ethically. Following the law, training workers, ke­eping records, and signing clear contracts cre­ates safer places. In addition, firms re­duce risks to workers and legal trouble­s. Caring about safety helps workers and projects as it shows the­ field acts properly. Moreover, it fosters a culture of re­sponsibility and benefits all involved, thus boosting the­ image of construction.

How to Calculate the Materials for Concrete Slab Construction

Reinforced concrete (RC) slabs are a fundamental component of reinforced concrete structures, providing a versatile platform for floors, building occupancy, and various structural elements. They are material-intensive structural elements that take a lot of human and material resources to construct. Accurately calculating the quantity of materials for a reinforced concrete slab is very important for cost estimation, material procurement, and efficient construction planning.

It is very pertinent to point out that the quantity of materials required for the construction of reinforced concrete slabs is dependent on the architectural and structural design requirements of the building. Large-spanning floors will require more materials to construct compared to shorter floor spans. Furthermore, special structural features such as cantilever slabs will demand more construction materials compared to simply supported or continuous floors.

The basic materials required for the construction of suspended floor slabs in a building are;

(1) Concrete (cement, sand, stones, and water)
(2) Reinforcement (floor mats and beam reinforcements)
(3) Formwork (temporary wooden support platform consisting of joists, plywoods/planks, and props).

This article discusses the key steps involved in the calculation of the quantity of these materials required for the successful execution of any building construction project. We will use the floor slab layout in Figure 1 as a case study. All the supporting beams are 230 x 400 mm beams.

Figure 1: Typical layout of a floor slab

Step 1: Determine Slab Geometry

The first step involves defining the slab’s geometric parameters. These include:

  • Length (L): The horizontal dimension of the slab measured in meters (m) or feet (ft).
  • Width (W): The perpendicular horizontal dimension of the slab measured in meters (m) or feet (ft).
  • Thickness (T): The vertical dimension of the slab measured in meters (m) or inches (in).

For the slab layout above,

Length (L) = 5000 + 5000 + 5000 + 230 = 15230 mm = 15.23 m
Width (W) = 6000 + 5000 + 230 = 11230 mm = 11.23 m
Thickness = 150 mm = 0.15 m

Step 2: Calculate Slab Volume

Once the geometry is defined, the volume (V) of the concrete required can be calculated using the formula:

V = L x W x T
V = 15.23 m × 11.23 m × 0.15 m = 25.654 m3

This volume represents the total amount of concrete needed to cast the slab.

Note: It’s important to ensure all units are consistent (e.g., all in meters or all in feet) to obtain accurate results.

Step 3: Account for Formwork

Formwork is the temporary structure that shapes and supports the concrete during the curing process. While not directly a concrete material, formwork influences the overall material quantity.

bamboo and timber joists are materials used in floor slab construction
Figure 2: Typical floor slab construction in a low-cost low-tech building construction

Two approaches can be considered:

  1. Direct Measurement: If the formwork design details are available, the volume of formwork materials (plywood, lumber, etc.) can be directly calculated using their dimensions and quantities.
  2. Percentage Addition: A common practice involves adding a percentage buffer to the concrete volume to account for formwork material absorption and potential overestimation errors. This percentage typically ranges from 5% to 10% depending on the formwork complexity and project tolerances.

For the project layout under consideration, the floor slab will supported by 250 mm x 3500 mm boards (planks) typically called 1″ x 12″ boards in Nigeria. Ideally, the width of the board is supposed to be 300 mm, but experience has shown that the boards available in local markets rarely meet the dimensional requirements. These boards will be supported by 2″ x 3″ softwood joists/stringers spaced at 600 mm c/c, which will be supported on bamboo props spaced at 600 mm c/c.

Wooden plank (sheathing) requirement

Area of floor slab = L x W = 15.23 m × 11.23 m = 171 m2
Area of wooden plank = b x h = 0.25 × 3.5 = 0.875 m2
Number of 1″ x 12″ planks required = 171/0.875 = 196 pcs

Making a 10% allowance for waste, damages, and offcuts = 1.1 × 196 = 216 pcs
(Note: This quantity does not account for the beam formwork requirements which should be calculated using the beam dimensions). If 1.2m x 2.4m marine boards are to be used, the quantity required will be approximately 60 pcs without accounting for waste and offcuts.

Boarding of a floor slab using 1" x 12" planks
Figure 3: Boarding of a floor slab using 1″ x 12″ planks

Side edge formwork
The thickness of the slab = 150 mm (therefore 1″ x 6″ board will be adequate for the edge formwork)
Perimeter of slab = 2(15230) + 2(11230) = 52920 mm
Number of 1″ x 6″ edge formwork required = 52920/3500 = 16 pcs

Timber Joist Requirement

To calculate the quantity of timber joists required, the recommended spacing of the joists is required. Using the length and width of the slab floor plan, you can calculate the total length of the timber joist required if the spacing is known. The number of wood materials required can then be obtained by dividing the total length of the timber joist by the supply length.

The supply length of most softwood timber beams in Nigeria is 12 ft (3.6m), but for quantification purposes, it is wise to use a shorter length, say 3.5 m. It is important to also ensure that good quality wood with the appropriate dimensions is supplied.

Recommended timber joist = 2″ x 3″ softwood (supply length is usually 3500 mm)
Spacing = 600 mm

Pick the length of the slab L = 15230 mm
Number of joists required along the length = 15230/3500 = 4.35 pcs
Number of rows of joists required along the width of the slab = (11230/600) + 1 = 20 rows

Therefore the number of 2″ x 3″ wood required = 20 × 4.35 = 87 pieces
Making a 10% allowance for waste, damages, and offcuts = 1.1 × 87 = 96 pcs

Alternatively, number of rows along the length (it can work with either the length or width) = (15230/600) + 1 = 27 rows
Length per row = 11230 mm
Total length of 2″ x 3″ required = 27 × 11230 = 303210 mm
Number of pieces required = 303210/3500 = 87 pieces

Figure 4: Typical joist and bamboo arrangement in a low-cost low-tech building construction

Bamboo prop requirements

To obtain the quantity of bamboo props required, the same steps can be followed as were done for the joists. However, we should know that the bamboo props are individual ‘posts’ standing on their own at a given spacing. Using the recommended spacing of the bamboo, it is possible to calculate the number of bamboo along the length of the floor slab and multiply it by the number of rows along the width of the slab. This should give the total quantity of bamboo required for the slab.

Spacing of bamboo props = 600 mm in all directions
Number of bamboo props along the length = (15230/600) + 1 = 27
Number of rows of bamboo along the width of the building = (11230/600) + 1 = 20 rows
Total number of bamboo props required = 27 × 20 = 540 pcs of bamboo.

If it is assumed that a minimum of 2 pcs of bamboo can be obtained per supply length of natural bamboo, then 270 lengths of fully matured bamboo stem should be ordered. It is also possible to calculate the quantity of bamboo props required ”room by room”.

Step 4: Estimate Material Quantities for Concrete Mix

The next step involves determining the quantities of individual components required to produce the concrete mix for the slab. This is based on the specified concrete mix design, typically denoted as “M XX” (where XX represents the characteristic compressive strength in MPa).

Common mix designs include M20, M25, and M30. For M20 concrete, a concrete mix of 1:2:4 with a maximum water-cement ratio of 0.5 should provide a minimum compressive strength of 20 N/mm2 after 28 days of curing.

Read…
How to achieve grade 25 concrete on-site

Material Quantities per Unit Volume of Concrete:

The specific quantities of cement, sand, and coarse aggregate (gravel) per unit volume of concrete depend on the mix design. However, a general guideline can be used for the initial estimation of a 1:2:4 mix ratio:

  • Cement: 6.5 bags of cement (50kg/bag) per 1 m3 of concrete
  • Sand: 1000 kg per 1 m3 of concrete
  • Coarse Aggregate: 1450 kg per 1 m3 of concrete

For 25.654 m3 of concrete;

  • Cement: (6.5 × 25.654 m3) = 167 bags of cement
  • Sand: (1000 x 25.654 m3) = 25654 kg of sand (25.6 tonnes of sand)
  • Coarse Aggregate: (1450 x 25.654 m3) = 37198 kg of granite (37.2 tonnes of granite)

Note: It is very important to consult the specific mix design for accurate material quantities. This information is typically provided by a structural engineer or obtained from ready-mix concrete suppliers.

Step 5: Quantify Reinforcement Steel

The amount of reinforcement steel required depends on the slab’s structural design. This information is typically provided on the structural drawings or by a structural engineer. The steel quantity is usually expressed in weight per unit area (kg/m²) or total weight for the entire slab (kg). It is therefore important that the accurate quantity of steel required is calculated from the structural drawing.

Common Steel Reinforcement Types:

  • Main bars: These primary bars resist the bending moments acting on the slab.
  • Distribution bars: These secondary bars distribute loads and prevent cracking.

Let us assume that for the slab above, Y12 @ 150 c/c has been recommended for the sagging and hogging areas, while Y10 @ 250 c/c has been recommended for the distribution bars. For 150 mm c/c spacing, a very rough estimate will show that 12.5 kg of 12mm bars is required per square metre of the slab, while 25% of the quantity of span reinforcement is required in the hogging areas. 15% of the span reinforcement is required as the distribution bar.

Figure 5: Typical rebar arrangement in a duplex construction

Read…
Quantity estimation guide for construction sites

Therefore, for the slab under consideration:

Bottom reinforcement required = 12.5 × (15.23 × 11.23) = 2137.9 kg of Y12 mm
Hogging reinforcement required = (25/100) × 2137.9 = 534.475 kg of Y12 mm
Distribution bar required = (15/100) × 2137.9 = 320.685 kg of Y10 mm

Alternatively;

Since the bottom reinforcement is placed at 150 mm c/c in both directions, we can calculate the total length of the rebars. We can increase the length and width of the rebars by 2m to account for the return bars. Ideally, this should extend by 0.15L into the span of the slab.

Total length of the slab = 15.23 + 2m = 17.23m
Number of rows required at 150 mm spacing = (11.23/0.15) + 1 = 76 rows
Total length required along the width of the slab = 17.23 × 76 = 1309.48 m

Total width of the slab = 11.23 + 2m = 13.23m
Number of rows required at 150 mm spacing = (15.23/0.15) + 1 = 103 rows
Total length required along the width of the slab = 13.23 × 103 = 1362.69 m

The total length of reinforcement required for the bottom rebars = 1309.48 + 1362.69 = 2672.17m

Rebars are supplied per 12m length, therefore, the total number of rebars required (in pieces)= 2671.17/12 = 223 lengths
Total number of bottom rebars required (in kg) = 2671.17 × 0.888 = 2372 kg of Y12 mm (Note that the unit weight of Y12mm bars is 0.888 kg/m)

Knowing that the top reinforcements in RC slabs are supposed to extend 0.3L into the span of the slab, the quantity of the top reinforcements can also be accurately determined.

Conclusion

Accurately estimating the material quantities for an RC slab is essential for efficient project planning, cost control, and material procurement. This article has outlined a comprehensive approach to achieve this goal. By systematically considering the slab geometry, formwork influence, concrete mix design, and reinforcement steel requirements, engineers and construction professionals can effectively determine the necessary materials.

It’s important to remember that the initial calculations may require further refinement based on project specifics and the involvement of a structural engineer. However, this process provides a valuable foundation for informed decision-making throughout the reinforced concrete slab construction process.

Rectangular vs Circular Columns: Strength, Design, and Uses

Columns are major structural members in buildings, with the sole purpose of transferring vertical and horizontal loads from beams and slabs to the foundation. The design of columns is very important in ensuring the stability and safety of buildings. The prevalent column geometries in building construction encompass rectangular, circular, and square cross-sections.

While rectangular and square sections are very popular in building construction due to the rectangular nature of walls, instances favouring circular or other column geometries do arise. For instance, for architectural reasons, standalone columns in the middle of halls or walkways are preferably circular due to aesthetic reasons. Furthermore, circular columns exhibit superior seismic performance compared to rectangular ones. A thorough understanding of these rationales is important for structural engineers and architects, particularly during the scheme development stage.

Circular column in a building

This knowledge facilitates informed decision-making regarding column geometry, thereby reducing the risk of inadequate construction work. It is to be emphasized that the selection of column geometry holds significant importance, similar to the determination of size, orientation, and positioning of both columns and beams within the structural system. While rectangular and circular columns satisfy the same structural functions, they offer distinct advantages and disadvantages in terms of strength, design considerations, and preferred applications.

Rectangular columns in an institutional building

Selection Criteria for Column Geometry in Building Construction

The configuration of a column’s cross-section, similar to other building elements, is governed by a multitude of factors such as architectural aesthetics, ease of construction, functional requirements, and structural demands.

Aesthetic Considerations

Architectural appeal is one of the prominent factors influencing column geometry. Square and rectangular columns can be easily placed within walls and covered with plastering, such that nobody will know that the columns were there in the first place. When square columns are not feasible due to structural size requirements, rectangular columns can be employed, such that the width will be equal to the width of the wall. This leads to a smooth flow and alignment of walls without undesirable projections.

Square columns can easily be hidden in walls

On the other hand, circular sections are often favoured for their inherent visual harmony. However, their application is frequently confined to specific contexts such as luxury mansions, monumental buildings, educational institutions, verandas, and public buildings. Circular columns are preferable when the column element is standing alone since they will rarely blend into walls without projecting out of the wall lines.

Formwork Considerations and Cost Implications

There is a significant disparity between the construction of column formwork of rectangular/square and circular sections. Achieving a smooth, curved profile for circular columns demands enhanced craftsmanship and superior formwork materials, often including new plywood or aluminium systems.

This complexity translates to challenges in attaining a high-quality concrete finish. Likewise, plastering a curved surface to the desired level is considerably more intricate compared to a flat plane, leading to increased construction time and cost. Given the budgetary constraints inherent in many medium-scale projects, clients often prioritize cost-effectiveness, making rectangular/square columns more favourable.

Wooden formwork construction for circular column
Wooden formwork construction for circular column

Furthermore, the pressure exerted on column formwork by fresh concrete can be more complicated when compared with square or rectangular sections. Consequently, stricter quality control measures and more elaborate formwork support systems are mandatory on-site. The increased level of workmanship required for these considerations translates to a significant cost increase compared to rectangular column construction.

Circular columns in a highrise building

Functional Requirements

The intended use of a space significantly influences column selection. In office environments, columns often serve as display surfaces for artwork, signage, or bulletin boards. Rectangular sections provide inherently planar surfaces ideal for such applications. This rationale extends to basement parking areas, where rectangular or square columns offer suitable flat surfaces for traffic flow notices, warnings, and signage – commonly observed in shopping malls and other commercial establishments.

Structural Considerations

While circular columns generally exhibit superior seismic performance, specific scenarios necessitate enhanced stiffness in a particular direction. This is particularly relevant for slender buildings with a limited base width and an extended length. To achieve this, engineers may favour a rectangular column with a larger dimension oriented towards the slender direction.

Moment of Inertia

A critical factor in column strength is the moment of inertia (I), a property that reflects a section’s resistance to bending. Circular columns boast a higher and more uniform I value across all axes compared to rectangular columns. This translates to superior resistance to bending moments and deflection, making them ideal for structures subjected to high lateral loads, such as bridges.

Shapes of rectangular and circular sections
Shapes of rectangular and circular sections

For rectangular sections, the moment of inertia (I) is given by;
I = bh3/12

For circular sections, the moment of inertia (I) is given by;
I = πD4/64

Let us consider a square column of dimensions 250 mm x 250 mm. This column will have an area of 0.0625 m2. A circular column that will produce a similar area will have a diameter of 282 mm.

The square section will have a moment of inertia I = bh3/12 = (0.25 × 0.253)/12 = 3.255 × 10-4 m4
The circular section will have a moment of inertia I = πD4/64 = (π × 0.2824)/64 = 3.104 × 10-4 m4

Therefore for sections of similar cross-sectional area, square sections have higher moment of inertia than circular sections.

Buckling Resistance

Another important aspect of strength consideration of columns is buckling, which is the tendency of a slender column to bend under compressive loads. Circular sections, due to their uniform distribution of material around the centroidal axis, offer superior buckling resistance compared to rectangular columns, especially when the load is not perfectly centred.

Material Efficiency

Rectangular columns, however, can be more material-efficient for specific loading conditions. By strategically orienting the rectangular section with the larger dimension towards the direction of higher bending moment, engineers can achieve optimal load-carrying capacity with less material compared to a circular column of equivalent area.

Reinforcement

Circular columns typically require more reinforcement bars due to their curved shape. While this can impact material costs, it also enhances their overall compressive strength.

Applications

Rectangular Columns:

  • Widely used in buildings due to their ease of construction and efficient space utilization in confined areas.
  • Preferred for load-bearing walls where columns can be integrated with the wall structure.
  • Suitable for situations where higher bending moment capacity is required in a specific direction by adjusting the rectangular section.

Circular Columns:

  • Ideal for bridge piers and other structures subjected to high lateral loads due to their superior bending and buckling resistance.
  • Used in open spaces or architectural features where their aesthetic appeal is valued.
  • Preferred in seismic zones due to their uniform distribution of strength across all axes.

Comparative Design of Square and Circular Columns

Investigate the design requirements of the two column cross-sections analysed above;
Square section = 250 x 250mm
Circular section = 282 mm (diameter)
Axial load = 1000 kN (No bending moment considered except secondary moments)
The effective length of the columns about the major and minor axis = 3000 mm√
Compressive strength of concrete = C20/25
Yield strength of reinforcement = 500 MPa

Square section design

Column slenderness about y-axis
Radius of gyration;   iy = h/√(12) = 7.2 cm
Slenderness ratio (5.8.3.2(1)); ly = l0y / iy = 41.6

Column slenderness about z-axis
Radius of gyration;  iz = b/√(12) = 7.2 cm
Slenderness ratio (5.8.3.2(1)); lz = l0z / iz = 41.6

Min end moment about y-axis; M01y = min(abs(Mtopy), abs(Mbtmy)) + eiyNEd = 7.5 kNm
Max end moment about y-axis; M02y = max(abs(Mtopy), abs(Mbtmy)) + eiyNEd = 7.5 kNm
Min end moment about z-axis; M01z = min(abs(Mtopz), abs(Mbtmz)) + eizNEd = 7.5 kNm
Max end moment about z-axis; M02z = max(abs(Mtopz), abs(Mbtmz)) + eizNEd = 7.5 kNm

Design bending moment about y-axis (cl. 5.8.8.2 & 6.1(4))
Nominal 2nd order moment; M2y = NEde2y = 14.3 kNm
M0ey = max(0.6M02y + 0.4M01y, 0.4M02y), max(0.6M02y – 0.4M01y, 0.4M02y)) = 7.5 kNm

Equivalent moment from frame analysis;                    
M0ey = max(0.6M02y + 0.4M01y, 0.4M02y) = 7.5 kNm

Design moment;                                                              
MEdy = max(M02y, M0ey + M2y, M01y + 0.5M2y, NEd × max(h/30, 20 mm))
MEdy = 21.8 kNm

Design bending moment about z-axis (cl. 5.8.8.2 & 6.1(4))
Nominal 2nd order moment; M2z = NEde2z = 13.4 kNm
M0ez = max(0.6M02z + 0.4M01z, 0.4M02z), max(0.6M02z – 0.4M01z, 0.4M02z)) = 7.5 kNm

Equivalent moment from frame analysis;                    
M0ez = max(0.6M02z + 0.4M01z, 0.4M02z) = 7.5; kNm

Design moment;                                                              
MEdz = max(M02z, M0ez + M2z, M01z + 0.5M2z, NEd × max(b/30, 20 mm))
MEdz = 20.9 kNm

Area of reinforcement provided = 6Y20 (Asprov = 1885 mm2)

Design axial resistance of section;  NRd = (Acfcd) + (Asfyd) = 1527.9 kN

Ratio of applied to resistance axial loads;   ratioN = NEd / NRd = 0.655
Exponent a = 1.46
Biaxial bending utilisation;  
UF = (MEdy / MRdy)a + (MEdz / MRdz)a = 0.899

DescriptionUnitProvidedRequiredUtilisationResult
Moment capacity (y)kNm34.2321.840.64PASS
Moment capacity (z)kNm40.4320.880.52PASS
Biaxial bending utilisation   0.90PASS

Circular section design

Column slenderness about both axis
Radius of gyration;   iy = iz = h / 4 = 7.1 cm
Slenderness ratio (5.8.3.2(1)); ly = lz = l0y / iy = 42.6

Ecc. due to geometric imperfections (y-axis); eiy = l0y /400 = 7.5 mm

Min end moment about y-axis; M01y = min(abs(Mtopy), abs(Mbtmy)) + eiyNEd = 7.5 kNm
Max end moment about y-axis; M02y = max(abs(Mtopy), abs(Mbtmy)) + eiyNEd = 7.5 kNm

Design bending moment about y-axis (cl. 5.8.8.2 & 6.1(4))
Nominal 2nd order moment;  M2y = NEde2y = 12.8 kNm
M0ey = max(0.6M02y + 0.4M01y, 0.4M02y), max(0.6M02y – 0.4M01y, 0.4M02y)) = 7.5 kNm

Equivalent moment from frame analysis;                    
M0ey = max(0.6M02y + 0.4M01y, 0.4M02y) = 7.5 kNm

Design moment;                                                              
MEdy = max(M02y, M0ey + M2y, M01y + 0.5M2y, NEd × max(h/30, 20 mm))
MEdy = 20.3 kNm

Design bending moment about z-axis (cl. 5.8.8.2 & 6.1(4))
Nominal 2nd order moment; M2z = NEde2z = 12.8 kNm
M0ez = max(0.6M02z + 0.4M01z, 0.4M02z), max(0.6M02z – 0.4M01z, 0.4M02z)) = 7.5 kNm

Equivalent moment from frame analysis;                    
M0ez = max(0.6M02z + 0.4M01z, 0.4M02z) = 7.5 kNm

Design moment;                                                              
MEdz = max(M02z, M0ez + M2z, M01z + 0.5M2z, NEd × max(b/30, 20 mm))
MEdz = 20.3 kNm

Resultant design moment; MEd = √(MEdy2 + MEdz2) = 28.7 kNm

Area of steel provided = 6Y20 (Asprov = 1885 mm2)

DescriptionUnitProvidedRequiredUtilisationResult
Moment capacity (y)kNm30.1120.310.67PASS
Moment capacity (z)kNm31.6820.310.64PASS
Combined capacitykNm30.1128.720.95PASS

As can be seen from the design results, under similar axial loading and support conditions, square columns and circular columns of equal area will demand approximately the same area of reinforcement. However, once uniaxial and bi-axial bending moments are involved, we should expect a completely different behaviour in the quantity of reinforcements required.

Conclusion

The choice between rectangular and circular columns depends on a project’s specific requirements. When prioritizing strength and buckling resistance for structures like bridges or seismic zones, circular columns are often favoured. However, for ease of construction, space optimization, and cost-effectiveness in buildings, rectangular columns are the preferred choice. Understanding the strengths and limitations of each shape allows engineers to make informed decisions for optimal structural performance and aesthetics.

In conclusion, the selection of column cross-sectional geometry necessitates a comprehensive evaluation involving architectural intent, functional considerations, and structural demands. While aesthetics may favour circular columns in specific scenarios, the complexities and cost implications associated with formwork often render rectangular/square sections the more pragmatic choice for many building projects.

Design of Piled Raft Foundations

Piled raft foundations represent an economical and practical solution for situations where a conventional raft foundation falls short of design requirements. This type of foundation system strategically integrates a limited number of piles beneath the raft, allowing the raft itself to still contribute significantly to the load-bearing capacity.

This key distinction separates them from traditional pile foundations, where the primary responsibility for supporting the structure rests solely on the piles. Consequently, piled raft foundations occupy a unique position within the broader category of pile foundation systems, necessitating more complex analytical, design, and application considerations.

The piled raft foundation system itself comprises three key components:

  1. Piles: These deep foundation elements transfer heavy structural loads to deeper and more stable soil layers.
  2. Raft: This shallow foundation element distributes loads across its footprint, primarily utilizing the near-surface bearing capacity of the soil.
  3. Supporting Soil: This plays a critical role in transmitting loads from the structure to the foundation system.
Load transfer in piled raft foundation systems
Figure 1: Load transfer in piled raft foundation systems

By combining the capabilities of piles and rafts, the piled raft foundation system offers a synergistic solution for supporting heavy structures. It leverages the deep load-bearing capacity of piles while simultaneously utilizing the shallow load-carrying capacity of the raft. This combined approach effectively resists both vertical and lateral loads, ensuring the serviceability and stability of the structure.

Load-settlement behaviour within piled raft systems is influenced by a multitude of factors. The physical and mechanical properties of the soil, raft foundation, and pile foundation all play a significant role. Furthermore, the construction sequence of the building itself can also exert an influence. While a range of simplified, approximate, and advanced methods exist for analyzing this complex system, the most realistic results are typically obtained through the use of advanced finite element analysis.

Piled raft foundations have become a prominent choice for a wide range of demanding construction projects. Their versatility and ability to handle challenging soil conditions make them ideal for high-rise buildings in urban centers, where maximizing footprint usage is crucial. Landmark structures like the Burj Khalifa in Dubai and the Shanghai Tower in China employed piled raft foundations to ensure stability and support their immense weight.

Figure 2: Burj Khalifa tower foundation was built with piled raft

Piled rafts also find application in infrastructure projects such as offshore wind turbine installations, where they provide a stable base for these towering structures amidst wind and wave loads. Furthermore, their ability to resist buoyancy is valuable for projects on sites with high water tables, such as waterfront developments or structures built on reclaimed land.

Types of Piled Raft Foundation

Piled raft foundation can be broadly categorised into two;

  • Piled raft for settlement control, and
  • Piled raft for load transfer

Piled raft for settlement reduction
While raft foundations can offer adequate bearing capacity, they may still be susceptible to excessive settlement. Traditionally, this issue is addressed by incorporating a basement and a basement raft, which effectively reduces the total load acting on the foundation system. However, when this approach is not feasible, an alternative solution involves introducing a limited number of piles beneath the raft.

These piles function by transferring a portion of the overall load away from the raft itself. As the piles do not need to carry the entire load, the required number is significantly lower compared to a traditional piled foundation design. Additionally, due to this load redistribution, the settlement experienced by the raft is brought within acceptable limits.

Piled raft for load transfer
The second category of piled rafts, designated as “conventional,” finds application in scenarios where the underlying soil exhibits pronounced weakness and a high water table is present. In such conditions, the adoption of a raft foundation becomes essential.

These rafts serve a dual purpose: first, resisting the buoyant forces exerted by the groundwater, and second, transmitting all net structural loads to the piles for transfer to deeper, more competent soil layers. Consequently, the number of piles necessitated in this scenario will be considerably greater compared to the previous case described.

In essence, piled raft foundations offer a synergistic approach to foundation design, leveraging the strengths of both piles and rafts. The raft’s ability to share the load and reduce differential settlements, coupled with the piles’ capacity to act as “stress reducers” and “settlement reducers” while enhancing the overall bearing capacity, paves the way for efficient and reliable foundation systems, particularly in challenging soil conditions.

Load Transfer and Sharing in Piled Raft Foundation

The load distribution in foundation systems plays a crucial role in determining the interaction between the structure and the underlying soil. While a footing or raft primarily affects the shallow soil layers (approximately 1-2 times its width), pile foundations transfer loads to deeper strata. Combining these two approaches in a combined Pile and Raft Foundation (CPRF) system creates a complex interplay influenced by several factors.

These factors include:

  • The rigidity of the raft: A stiffer raft tends to distribute load more evenly across the foundation, while a flexible raft allows for greater load transfer to the piles.
  • Soil properties: The stiffness and bearing capacity of the underlying soil layers significantly impact the load transfer mechanisms.
  • Pile characteristics: The number, depth, and rigidity of the piles within the raft influence how the load is shared between the piles and the raft.

The piled raft foundation system, responsible for transferring a structure’s load to the underlying soil, presents a complex interaction that have captivated researchers for years. Early notions, often overly conservative, assumed that the raft, in direct contact with the soil, offered no resistance to applied loads.

However, recent research works challenge this assumption, highlighting the raft’s significant contribution, particularly in clayey soils subjected to substantial structural loads. These studies reveal that the raft bears a portion of the load, while the piles carry the remaining portion through a creep mechanism.

The analysis and design of piled raft foundations, therefore, demand careful consideration of various critical factors, including:

  • Raft thickness and dimensions
  • Pile length, diameter, and configuration within the raft
  • Underlying soil properties
  • Stiffness characteristics of both the pile and raft

These factors collectively influence the load-sharing mechanism between the piles and the raft, ultimately impacting the stability and serviceability of the structure.

Furthermore, the success of piled raft foundation hinges on understanding the two key interaction types:

  • Pile-to-pile interaction: This interaction depends heavily on the soil’s elastic modulus, the pile slenderness ratio (s/d), and the pile length. Ignoring this interaction can lead to underestimating settlements and bending moments in the raft, compromising structural safety.
  • Pile-to-raft interaction: This interaction influences the load distribution between the piles and the raft, affecting the overall settlement behaviour of the foundation system.

Nevertheless, combined pile and raft foundation systems have challenged researchers for years due to the complexities associated with load-sharing and analysis. Several researchers have endeavoured to address this challenge, notably, by proposing simplified methods that incorporate various simplifications. However, these methods should be employed with caution due to their inherent limitations.

Quantifying the load contribution of each element within a piled raft system remains a topic of ongoing investigation. While some researchers suggest piles carry 50-80% of the total load, others provide a wider range of 30-60% for the raft’s contribution, emphasizing the dependence on factors like soil conditions, pile length, and spacing. Their research also highlights a decreasing raft contribution with denser pile spacing and increased pile length.

Further research strengthens the argument for the raft’s significant role, attributing up to 50% of the structural load to its contribution.

Classical Methods of Detemining Pile-Raft Load

Some classical methods of detemining pile-raft load sharing are discussed below.

Randolph Method
In the method proposed by Randolph (1994), load sharing ratio between pile group and pile raft, pile raft stiffness and settlement of piled raft can be calculated by using Eq. 1 to Eq. 4.

α = QR/QPG = 0.2/[(1 – 0.8(KR/KPG)] × KR/KPG ——— (1)
β = QR/QPG = 1/(1 + α) ——— (2)
KPR = [1 – 0.6(KR/KPG)]/[1 – 0.64(KR/KPG)] × KPG ——— (3)
S = Q/KPR ——— (4)

Where,
α = Load sharing ratio between raft and pile group,
β = Load sharing ratio between pile group and pile raft,
KR = Stiffness of the raft
KPG = stiffness of the pile group
KPR = stiffness of the piled raft
S = Settlement of piled raft,
Q = Design load

Poulos-Davis-Randolph (PDR) Method
In the method proposed by Poulos, Davis, and Randolph, load sharing ratio between raft and piled raft (X) can be determined using Randolph (1994) method and piled raft settlement (S) can be established using Poulos and Davis (1980) method.

X = Qr/Qpr = [(1 – αrp)kr]/[kpg + (1 – 2αrp)kr] ——— (5)
S = Qpr/kpr ——— (6)
kpr = [kpg + (1 – 2αrp)kr]/[1 – αrp2 (kr/kpg)] ——— (7)
αrp = 1 – [In(rc/r)/ς] ——— (8)
rc = √(A/nπ) ——— (9)
ς = In(rm/r) ——— (10)
rm = {0.25 + ξ[2.5ρ(1 – v) – 0.25]}L ——— (11)
ρ = Gavg/Gl ——— (12)
ξ = Gl/Gb ——— (13)

where;
Qr = Load carried by the raft
Qpr = Load carried by piled raft,
kr = stiffness of the raft
kpg = stiffness of the pile group
kpr = Stiffness of the piled raft,
αrp = Interaction factor,
A = Raft area,
n = Number of piles
L = length of piles
rm = Maximum radius from pile axis,
Gl = Shear modulus of soil along pile shaft
Gb = Shear modulus of soil at pile end
ν = Poisson’s ratio of soil.

Design of Piled Raft

According to Poulos (2001), the design of a piled raft foundation can be effectively divided into three distinct stages. The initial stage focuses on a preliminary analysis, estimating the impact of varying pile numbers on the overall load capacity and settlement of the structure. This analysis is typically approximate in nature.

Additionally, preliminary design stages often benefit from incorporating load-sharing ratios and settlement values derived from empirical studies and case histories. The expertise of designers familiar with piled raft systems remains an important component in achieving optimal outcomes.

The second stage goes deeper, aiming to identify the specific locations where piles are necessary and providing an initial indication of the required piling specifications.

Finally, the third stage represents the detailed design phase. Here, a more refined analysis is employed to confirm the optimal number and positioning of the piles. Additionally, this stage gathers crucial information for the structural design of the entire foundation system.

Complexities inherent to piled raft systems necessitate the use of sophisticated analytical methods during the design stage. These methods, such as the finite element method (FEM), boundary element method (BEM), equivalent element method (EEM), and plate-on-spring method (POSM), account for the numerous variables influencing the system’s behaviour.

Approximate methods like the “strip on springs” approach and the “plate on springs” offer further avenues for analysis. These methods offer simplified representations of the raft and piles (as springs) to understand their interaction.

For more detailed analysis, researchers recommend resorting to numerical methods, with the Finite Element Method (FEM) being the most prevalent choice. Software like SAP2000 and PLAXIS 3D foundation are prime examples of FEM-based solutions. FEM provides approximate solutions for various nonlinear engineering problems, including those encountered in combined piled raft foundation analysis.

It is important to note that while simplified methods can provide reasonable results for preliminary assessments, numerical methods like FEM offer a superior level of accuracy and detail for complex piled raft foundation systems. Recognizing the limitations of each approach is paramount for selecting the most appropriate analytical tool for a specific project.

Summarily, understanding the load-sharing mechanisms within piled raft foundation remains a dynamic field of research. While simplified methods offer initial insights, numerical methods like FEM provide a more robust and accurate means for analyzing these complex foundation systems. Choosing the appropriate analytical approach requires careful consideration of the project’s specific needs and complexities.

Preliminary Design Example of Piled Raft Foundation System

This section considers the preliminary design of a proposed piled raft foundation system. The raft is 750 mm thick and the superstructure load distribution on the raft is shown in Fgure 3. The initial proposed distribution of the piles, comprising of 9 number of 600 mm diameter piles is shown in Figure 4.

Figure 3: Load distribution on the raft foundation system
Figure 4: Preliminary arrangement of the piled raft system

Thickness of raft = 750 mm
Modulus of subgrade reaction ks = 10000 kN/m2/m
Modulus of horizontal compressibility nh (medium dense wet sand) = 4000 kN/m2/m
Pile diameter = 600 mm
Depth of pile = 10 m

Horizontal modulus of subgrade reaction = nh(z/d) ——— (14)

The horizontal modulus of subgrade reaction was used in modelling the piles, and the spring stiffness varied with depth according to equation (14).

3D render of the piled raft foundation on Staad Pro software
Figure 5: 3D render of the piled raft foundation on Staad Pro software
Finite element model of the piled raft foundation on Staad Pro software
Figure 6: Finite element model of the piled raft foundation on Staad Pro software
Figure 7: Bending moment diagram of the piles
Figure 8: Load applied on the foundation

Total load applied on the foundation = 4(350) + 2(500) + 6(600) = 6000 kN

From the analysis results,
Total load transferred to the piles = 900 kN
Therefore, total load resisted by the raft = 6000 – 900 = 5100 kN

In this case, about 85% of the load is resisted by the raft foundation. If it is a piled raft foundation where the piles are to be used in load transfer, the arrangement of the piles will have to be changed. However, if it is a system where the piles are to be used for settlement control, the pile arrangement can be evaluated for acceptance or rejection.

In a different scenario when the number of piles was increased to 18 (additional piles were introduced along the column gridlines), 69% of the load was resisted by the raft. Therefore, preliminary analysis requires a careful consideration of the location and number of piles in the system.

Conclusion

Piled raft foundation systems offer a powerful solution for navigating complex soil conditions and supporting substantial loads. The design process involves a meticulous three-stage approach, starting with a preliminary analysis, then progressing to detailed location and quantity determination of piles, and finally culminating in a refined analysis for optimal pile placement and structural design of the entire foundation.

This staged approach ensures an efficient and cost-effective foundation that leverages the strengths of both raft foundations and pile foundations. Piled raft systems are a versatile solution for high-rise buildings, infrastructure projects, and construction on challenging sites, providing the stability and support necessary for a wide range of demanding applications.

Sources and Citations

Randolph M. F. (1994). Design methods for pile groups and piled rafts, 13th ICSMFE, New Delhi, India, 61-82.
Poulos H.G. and Davis E.H. (1980). Pile foundation analysis and design, John Willey and Sons, New York, USA.
Poulos H.G. (2001). Piled raft foundations: designs and applications. Geotechnique 51(2):95-113

A General Overview of Dynamics of Structures

Structural dynamics is a field of study that discusses the behaviour of structures subjected to dynamic loads. It encompasses the analysis, design, and evaluation of structures under the influence of various forces and vibrations. While static analysis focuses on the behaviour of structures under constant or slowly varying loads, dynamics of structures explores how structures respond to dynamic loads, which are forces that change rapidly with time.

In addition to static loads, a structural system can be subjected to variable (dynamic) loads induced by factors such as wind and wave action, earthquakes, impact, blasts, and vehicular/pedestrian traffic (which causes vibration and fatigue in bridges). Therefore, understanding the dynamic behaviour of structures is important for ensuring the safety and serviceability of structures in scenarios involving:

  • Earthquakes: Ground motions induced by earthquakes can cause significant dynamic forces on structures, potentially leading to failure if not properly accounted for in design.
  • Wind: Wind loads can create significant dynamic effects, especially on slender structures like tall buildings and suspension bridges.
  • Vibrations: Structures subjected to human activity, machinery operation, or traffic can experience vibrations, which can lead to fatigue, discomfort, or even damage if not managed effectively.
  • Blast loads: Explosions and other rapid pressure changes can create extremely dynamic forces that need to be considered in the design of structures in specific environments.

This article provides a comprehensive exploration of the key concepts and methodologies involved in the analysis and design of structures subjected to dynamic loads.

Dynamics of buildings
Figure 1: Typical Dynamics of buildings

Dynamic Analysis

The methods of analysis used for static loads are insufficient to analyze the ‘dynamic’ or ‘time-varying’ loads and their impacts. When compared to the values of displacement that are produced by static loading, the values that are produced by the response of structural members to time-varying loads will likewise be time-varying, and this can result in substantially larger values.

To make the concept of structural dynamics clearer, let us consider a structural element that is subjected to an externally applied load. By considering the equilibrium of applied forces and the internal forces that correspond to those forces, it is always possible to compute the internal stresses and displacements of a structure, regardless of whether the force that is being applied is “static” or “dynamic.”

Assuming that the structure is linearly elastic, the internal forces and the displacements are linearly proportional. If, on the other hand, the force is applied in a dynamic manner, two additional types of internal forces are generated as a consequence. The first of these is referred to as the “inertia forces,” and it is related to the acceleration. The second of these is referred to as the “damping forces,” and it is proportional to the velocity.

In this article, we are going to present a fundamental introduction to the principles of structural dynamics, and how it can be extended to the design of structures.

Importance of Structural Dynamics Analysis

A comprehensive dynamic analysis of structures can reveal the potential for serviceability failures that would be entirely undetectable through a purely static evaluation. For instance, there have been documented cases of oil rigs being decommissioned in relatively calm seas due to the initiation of oscillations that were unacceptably uncomfortable for the crew.

Similarly, electric transmission lines have been known to develop severe dynamic oscillations, referred to as “galloping,” to the extent that the lines made contact. While this phenomenon may not necessarily lead to structural collapse, it undoubtedly constitutes a serviceability failure from the perspective of electricity consumers.

Structures under construction are especially susceptible to dynamic effects. For example, temporary damping measures were deemed necessary for the towers of the Forth Road Bridge in Scotland to mitigate dynamic effects before the installation of the main cables. Even from a purely structural strength perspective, dynamic analysis can be crucial if fatigue is a primary concern.

In such scenarios, it becomes essential to predict not only the magnitude of stresses within the structure but also the frequency at which various stress levels occur. This is because a consistently applied low stress can have a more detrimental fatigue impact than an occasional instance of higher stress.

Characteristics of a Structural Dynamic Problem

A structural dynamic problem differs from a static loading problem in two significant ways. Firstly, the dynamic problem is characterized by its time-varying nature. Since both the loading and the response change over time, a dynamic problem does not have a single solution like a static problem does. Instead, the analyst must determine a series of solutions corresponding to different times of interest in the response history. As a result, dynamic analysis is inherently more complex and time-consuming than static analysis.

The second and more fundamental distinction between static and dynamic problems is illustrated in Figure 2. When a simple beam is subjected to a static load p (as shown in Figure 2a), its internal moments, shears, and deflected shape depend solely on this load and can be calculated using established principles of force equilibrium.

Figure 2: Basic difference between static and dynamic loads: (a) static loading; (b) dynamic loading.

However, when the load p(t) is applied dynamically (as shown in Figure 2b), the resulting beam displacements depend not only on this load but also on inertial forces that resist the accelerations causing them. Consequently, the internal moments and shears in the beam must balance not only the externally applied force p(t) but also the inertial forces resulting from the beam’s accelerations.

Inertial forces, which oppose the accelerations of the structure, are the key distinguishing characteristic of a structural dynamics problem. Generally, if the inertial forces constitute a significant portion of the total load equilibrated by the internal elastic forces of the structure, the dynamic nature of the problem must be considered in its solution.

On the other hand, if the motions are so slow that the inertial forces are negligible, the response analysis for any specific time can be conducted using static structural analysis methods, despite the load and response being time-varying.

Equations of Motion

The mass, stiffness, and damping (energy absorption capability), of a linearly elastic structural system are the basic physical parameters that define the system when it is subjected to external dynamic loading. Consider the ‘dash-pot’ model (representing a simple building with a single storey) that is presented in Figure 3. This model can be used to demonstrate the fundamental idea behind dynamic analysis.

Figure 3: Vibration modelling of a single storey structure

The structure is subjected to a time-varying force denoted by f(t), in which k is the spring constant that links the lateral storey deflection (x) to the storey shear force, and c is a damping coefficient that relates the dashpot’s damping force to the velocity. If it is assumed that all of the mass, m, is located at the beam, then the structure will be considered a single-degree-of-freedom (SDOF) system.

It is possible to write the equation of motion of the system as follows;

mẍ + cẋ + kx = f(t) ——– (1)

Types of Vibration

Free Vibration

While our initial discussion addressed the impact of time-varying loads on structural behaviour (dynamic behaviour), a foundational understanding of vibration in simple structures, independent of dynamic loads, proves most beneficial. This specific type of vibration, termed “free vibration,” arises whenever a structure experiences a disturbance from its state of static equilibrium. The initiation of free vibrations can be attributed to either impulsive events such as a collision or explosion, or to sudden movements in the structure’s support system.

In this case, the system is set to motion and allowed to vibrate in the absence of applied force f(t). Letting f(t) = 0, equation (1) becomes:

mẍ + cẋ + kx = 0 ——– (2)

Dividing equation (2) by the mass m, we have:
ẍ + 2ξωẋ + ω2x = 0 ——– (3)

Where;
2ξω = c/m ——– (4)
ω2 = k/m ——– (5)

The solution to the equation depends on whether the vibration is damped or undamped.

Undamped Free Vibration

In the absence of not only time-dependent forces, but also any mechanisms for energy dissipation within the vibrating system, the resulting motion can be classified as both free and undamped. Realistically, energy losses due to factors such as friction and air resistance are unavoidable. Therefore, the concept of undamped vibration, while theoretically useful, represents an idealized scenario that disregards these energy-dissipating phenomena. Nevertheless, it remains a valuable tool for theoretical analysis.

In this case, c = 0, and the solution to the equation of motion may be written as:
x = Asinωt + Bcosωt ——– (6)

where ω = √(k/m) is the circular frequency. A and B are constants that can be determined by the initial boundary conditions.

Undamped free vibration
Figure 4: Undamped free vibration

Damped Free Vibration

The phenomenon of damping arises from the inevitable energy loss that occurs during vibration. This lost energy is either dissipated as heat within the structure or radiated outwards, often in the form of sound waves.

Internal friction within the structural materials themselves contributes a portion to this energy loss, with frictional losses at structural joints playing an additional role. While air resistance can also contribute to energy dissipation, it is typically considered a secondary factor.

To model the effects of damping in a simplified manner, engineers often employ a theoretical element known as a “dashpot” system.

If the system is not subjected to applied force and damping is present, the corresponding solution becomes:
x = A exp(λ1t) + B exp(λ2t) ——– (7)

λ1 = ω[-ξ + √(ξ2 – 1)] ——– (8)
λ2 = ω[-ξ – √(ξ2 – 1)] ——– (9)

The solution of equation (7) changes its form with the value defined as:

ξ = c/2√mk ——– (10)

Damped free vibration
Figure 5: Damped free vibration

Forced Vibration

When a structure experiences time-varying loads or continuous disturbances to its supports, the resulting motion is classified as forced vibration. The specific time-dependent influence that triggers this motion is termed excitation. The nature of the forced vibration – its frequency, amplitude, and overall behaviour – is directly tied to the characteristics of the excitation itself.

In essence, the excitation acts as an external “driving force” that dictates the response of the structure. This response can vary significantly depending on the excitation. For instance, a harmonic excitation (a smoothly oscillating force) will lead to a harmonic vibration with the same frequency but potentially a different amplitude. Conversely, a more impulsive excitation, like a sudden impact, can induce a transient vibration with a complex frequency spectrum.

If a structure is subjected to a sinusoidal motion such as a ground acceleration of = F sinωft, it will oscillate and after some time the motion of the structure will reach a steady state. For example, the equation of motion due to the ground acceleration (from equation (3)) is:

ẍ + 2ξωẋ + ω2x = Fsinωft ——– (11)

The solution to the equation we’ve been examining can be broken down into two key components. The first, known as the complementary solution (represented by equation 6), captures the transient behaviour of the system. If the system experiences any damping, the oscillations associated with this component will gradually diminish over time.

This decay effect eventually leads the system to reach a steady state, where it vibrates with a constant amplitude and frequency. This sustained vibration, termed forced vibration, is solely described by the second part of the solution, the particular solution, expressed as:

x = C1sinωft + C2cosωft ——– (12)

A key observation here is that the forced vibration occurs at the frequency of the excitation force, denoted by ωf, rather than the natural frequency of the structure itself, ω. Essentially, the external force dictates the frequency of the vibration. The term -F/ω² within the particular solution represents the static displacement D caused by the force, essentially accounting for the inertia of the structure.

Now, let’s explore the dynamic response of the structure under varying excitation frequencies relative to its natural frequency (ω):

  • Low-Frequency Excitation (ωf/ω > 1): When the applied force oscillates at a frequency significantly lower than the structure’s natural frequency, the response exhibits a characteristic termed quasi-static. In this regime, the system behaves as if it were under a constant load. The response is primarily governed by the stiffness of the structure, and the resulting displacement amplitude closely resembles the static deflection that would occur under a constant force of the same magnitude.
  • High-Frequency Excitation (ωf/ω < 1): Conversely, when the excitation frequency is much higher than the natural frequency, the response becomes primarily dependent on the mass of the structure. The displacement amplitude in this case is generally less than the static deflection (D < 1). This is because the structure’s inertia can effectively resist the rapidly oscillating force.
  • Resonance (ωf/ω ≈ 1): The most critical scenario arises when the excitation frequency nears the natural frequency of the structure (ωf/ω ≈ 1). Under these conditions, a phenomenon known as resonance occurs. Resonance drastically amplifies the displacement amplitude, potentially leading to catastrophic consequences for the structure. In essence, the external force synchronizes with the structure’s natural tendency to vibrate, causing a dramatic buildup of energy within the system.

The simplest periodic motion equation can be written as;

y(t) = Asin(ωt + φ0) ——– (13)

where A is the amplitude of vibration, φ0 is the initial phase of vibration, and t is time. This case is presented in Fig. 6a. The initial displacement y0 = Asinφ0 is measured from the static equilibrium position. The number of cycles of oscillation during 2π seconds is referred to as circular (angular or natural) frequency of vibration ω = 2 π/T (radians per second or s-1), T (s) is the period of vibration. Figure 6b, c presents the damped and increased vibration with constant period.

Figure 6: Types of oscillatory motions

Degrees of Freedom

The concept of degrees of freedom (DOF) plays a crucial role in both statics and structural dynamics. While the definition remains the same – the number of independent parameters that uniquely define the spatial positions of all points in a structure – its interpretation differs subtly between these two fields.

In statics, the DOF is often associated with structures modeled as collections of absolutely rigid discs. Here, a DOF greater than or equal to one signifies a geometrically changeable system. Such a system wouldn’t typically be considered a realistic engineering structure, as real structures exhibit some level of deformation. Conversely, a DOF of zero implies a geometrically unchangeable and statically determinate system – a structure with a unique solution for its equilibrium under applied loads.

However, in structural dynamics, the focus shifts to the deformation of the structural members themselves. A DOF of zero in this context indicates an absolutely rigid body, incapable of any displacement in space. This scenario is purely theoretical, as all real structures exhibit some degree of flexibility.

Furthermore, structures can be broadly classified into two categories based on their DOF:

  1. Structures with Concentrated Parameters: These represent structures where the distributed mass of individual members can be neglected compared to lumped masses concentrated at specific points along the members.
  2. Structures with Distributed Parameters: These structures are characterized by a uniform or non-uniform distribution of mass throughout their components. Analyzing these structures often requires more complex mathematical tools compared to those used for concentrated parameter systems.

From mathematical point of view, the difference between the two types of systems is the following: the systems of the first class are described by ordinary differential equations, while the systems of the second class are described by partial differential equations.

Distributed Mass Systems

While the lumped mass model offers a valuable simplification for many structures, it’s important to recognize that all real structures are fundamentally distributed mass systems. This implies that they can be conceptually divided into an infinite number of infinitesimal particles. As a consequence, if a distributed mass system experiences repetitive motion, it theoretically possesses an infinite number of natural frequencies and corresponding mode shapes – unique vibration patterns associated with each frequency.

However, the seemingly overwhelming complexity of analyzing a distributed system can be effectively bridged once its natural frequencies and mode shapes are determined. At this point, the analysis becomes mathematically equivalent to that of a discrete system, where the structure’s behavior is represented by a finite number of lumped masses interconnected by springs or other idealized elements.

The key lies in recognizing that, in practical scenarios, only a limited number of modes, typically those associated with lower frequencies, significantly contribute to the overall dynamic response of the structure. By focusing on these dominant modes, engineers can effectively convert the problem of a distributed mass system into a more manageable discrete system. This approach allows for accurate analysis using computationally efficient methods, enabling engineers to assess the dynamic behavior of real-world structures without getting bogged down by the theoretical infinite nature of distributed systems.

Conclusion

Dynamics of structures considers the interplay between time-varying external forces, internal resistance, and the inherent flexibility of structures. This article has looked into the fundamental concepts of free and forced vibrations, recognizing the crucial role of natural frequencies and damping in shaping a structure’s response.

However, real-world forces and ground motions can be incredibly complex. To accurately predict a structure’s behavior under these conditions, engineers typically rely on numerical analysis techniques. One of the most prevalent methods for solving such complex problems is the finite element method.

The analysis of structures subjected to dynamic loads hinges on the ability to model their behaviour effectively. While lumped mass systems offer a practical approach for many structures, the underlying reality of distributed mass systems with infinite natural frequencies cannot be ignored. The key lies in identifying the dominant modes that significantly influence the dynamic response, allowing us to transform the seemingly intractable distributed system into a more manageable discrete one.

In essence, the lumped mass model serves as a powerful tool for approximating the behavior of complex distributed systems. By strategically selecting the most influential modes, engineers can achieve a high degree of accuracy while maintaining computational tractability. This balance between theoretical completeness and practical feasibility is crucial for ensuring the safety and performance of structures subjected to dynamic loads.

Understanding the dynamics of structures equips engineers with the knowledge to design and build resilient structures that can withstand the challenges of the real world. From earthquakes and windstorms to traffic vibrations and human activity, structures must be able to withstand the complex effects of time-dependent loads without compromising safety, functionality, or serviceability. By mastering the principles of dynamics of structures, engineers can ensure that these structures perform their intended function in harmony with the dynamic forces that surround them.

Bending Moment Diagrams for Frames

In the field of structural engineering, understanding the internal forces acting on framed structures is important for the design of such structures. Among these forces, bending moment plays a very important role in influencing the behaviour of beams and columns in framed structures under various loading conditions. This article discusses the concept of bending moment and its visualization through bending moment diagrams (BMDs) for framed structures.

Understanding Bending Moment

Imagine a beam supported at its ends and subjected to a transverse load (a load acting perpendicular to the beam’s axis). This load induces internal forces within the beam, causing it to bend. The bending moment at any point along the beam’s length represents the turning effect (rotational tendency) or moment created by the internal forces acting on that specific section. It is essentially the product of the force (F) acting at a perpendicular distance (d) from the point of interest, expressed mathematically as:

M = F × d

The bending moment tends to rotate the beam section about an axis perpendicular to its longitudinal axis. A positive bending moment signifies concavity downwards while a negative bending moment indicates concavity upwards.

Bending Moment Diagrams

A bending moment diagram (BMD) is a graphical representation of the bending moment throughout the length of a beam or a member in a framed structure. This diagram helps visualize the variation of the bending moment along the member, enabling engineers to identify critical sections where the moment is highest and assess the potential for bending failure. Bending moment diagrams are plotted in the tension zone of structures.

With the advent of numerous structural analysis and design software, bending moment diagrams can easily be generated using results from finite element analysis.

Steps to Construct a Bending Moment Diagram

  1. Determine the support reactions: This involves analyzing the entire frame to calculate the forces acting at the supports due to the applied loads. For statically determinate frames, the equations of equilibrium are sufficient for determining the support reactions but for statically indeterminate structures, methods like the force method can be used.
  2. Cut the member: Imagine isolating a specific section of the member by making a virtual cut at a chosen point.
  3. Treat the section as a free body: Draw a free-body diagram of the isolated section, including all external forces (support reactions and applied loads) acting on it.
  4. Apply equilibrium equations: Utilize the principles of equilibrium (summation of forces and moments equal to zero) to solve for the internal shear force (V) and bending moment (M) at the cut section.
  5. Repeat for different sections: Choose multiple points along the member’s length and repeat steps 2-4 to determine the shear force and bending moment at each point.
  6. Plot the values: Plot the calculated bending moments on the vertical axis and the member’s length on the horizontal axis, connecting the points to form a smooth curve. This curve represents the bending moment diagram for the member.

Interpreting Bending Moment Diagrams

Bending moment diagrams reveal valuable information about the bending behaviour of a framed structure:

  • Zero bending moment: Points on the BMD where the curve crosses the horizontal axis indicate locations where the bending moment is zero. These points typically occur at supports or points of contraflexure.
  • Maximum and minimum bending moment: The peak positive and negative values on the BMD represent the sections experiencing the highest and lowest bending moments, respectively. These sections are often critical for design considerations.
  • Slope of the BMD: The slope of the BMD at any point signifies the rate of change of the bending moment. A positive slope indicates an increasing moment, while a negative slope represents a decreasing moment.

Applications of Bending Moment Diagrams

Bending moment diagrams are instrumental in various aspects of structural engineering, including:

  • Structural design: They aid in selecting appropriate beam sizes and materials by identifying sections with high bending moments, ensuring sufficient strength and preventing failure.
  • Deflection analysis: By knowing the bending moment distribution, engineers can estimate the deflection of the frame using various methods, evaluating its serviceability under load.
  • Reinforcement detailing: In reinforced concrete structures, BMDs guide the placement of steel reinforcement to counteract the bending moment and ensure adequate structural capacity.

Typical Bending Moment Diagrams for Rigid Frames

F = Total Load
IAB = ICD (the moment of inertia of the columns are equal)
K =IBCh/IABL
k1 = K + 2
k2 = 6K + 1
k3 = 2K + 3
k4 = 3K + 1

Rigid frame subjected to gravity uniformly distributed load on the beam

Bending moment diagram of a frame subjected to gravity uniformly distributed load on the beam (fixed support)
Bending moment diagram of a frame subjected to gravity uniformly distributed load on the beam (fixed supports)
FOR FIXED SUPPORTS
HA = HD = Fl/4hk1
VA = VD = F/2
MA = MD = Fl/12k1
MB = MC = Fl/6k1
Bending moment diagram of a frame subjected to gravity uniformly distributed load on the beam (pinned supports)
FOR PINNED SUPPORTS
HA = HD = Fl/4hk3
VA = VD = F/2
MA = MD = 0
MB = MC = HAh = Fl/4k3

Rigid frame subjected to a point load on the beam

Bending moment diagram of a frame subjected to a point load on the midspan of the beam (fixed support)
Bending moment diagram of a frame subjected to a point load on the midspan of the beam (fixed supports)
FOR FIXED SUPPORTS
HA = HD = 3Fl/8hk1
VA = VD = F/2
MA = MD = Fl/8k1
MB = MC = Fl/4k1
Bending moment diagram of a frame subjected to a point load on the midspan of the beam (pinned supports)
Bending moment diagram of a frame subjected to a point load on the midspan of the beam (pinned supports)
FOR PINNED SUPPORTS
HA = HD = 3Fl/8hk3
VA = VD = F/2
MA = MD = 0
MB = MC = HAh = 3Fl/8k3

Rigid frame subjected to a horizontal uniformly distributed load on the column

Bending moment diagram of a rigid frame subjected to horizontal uniformly distributed load on the column
Bending moment diagram of a rigid frame subjected to horizontal uniformly distributed load on the column (fixed supports)
FOR FIXED SUPPORTS
HA = F - HD
HD = Fk3/8k1
VA = -FhK/lk2 = -VB
MA = Fh/4[(K + 3)/6k1 + (4K + 1)/k2]
MB = h(HA - ½F) - MA
MC = HDh - MD
MD = Fh/4[(K + 3)/6k1 - (4K + 1)/k2]
Bending moment diagram of a rigid frame subjected to horizontal uniformly distributed load on the column (pinned supports)
Bending moment diagram of a rigid frame subjected to horizontal uniformly distributed load on the column (pinned supports)
FOR PINNED SUPPORTS
HA = F/8[(6k3 - K)/k3]
HD = F - HA
VD = -VA = Fh/2l
MA = MD = 0
MB = h(½F - HD) = 3Fhk1/8k3
MC = HDh = Fh/8[(2k3 + K)/k3]

Rigid frame subjected to a horizontal point load at the top

Bending moment diagram of a rigid frame subjected to horizontal point load at the top
FOR FIXED SUPPORTS
HA = HD = F/2
VA = -VD = -3FhK/Lk2
MA = MD = Fhk4/2k2
MB = MC = 3FhK/2k2
Bending moment diagram of a rigid frame subjected to horizontal point load at the top (pinned supports)
Bending moment diagram of a rigid frame subjected to horizontal point load at the top (pinned supports)
FOR PINNED SUPPORTS
HA = HD = F/2
VD = -VA = Fh/l
MA = MD = 0
MB = MC = Fh/2

Conclusion

Understanding bending moments and their visualization through bending moment diagrams is fundamental for structural engineers. By mastering this concept, engineers can effectively analyze framed structures, optimize designs, and ensure the safety and serviceability of their creations under various loading conditions.

Wind Load Analysis of Tank Farms and Other Cylindrical Structures

Cylindrical structures like tanks, silos, and chimneys are widely used in various industries, including oil and gas, agriculture, chemical processing, and water storage. These structures are susceptible to wind loads, which can cause significant stresses and potential failures if not properly analyzed and designed.

In industrial and agricultural settings, cylindrical above-ground vertical tank farms are commonly used for the storage of various liquids like petroleum, oil, water and fuel. These tanks are typically welded, thin-walled structures with large diameters, making them susceptible to buckling under wind loads when empty or partially filled.

The failure of such tanks can have devastating consequences, often resulting in significant financial and human losses. Additionally, these failures pose a serious threat to public safety and can have a detrimental impact on the environment.

Cylindrical structures can be affected by wind load
Cylindrical structures can be affected by wind load

This article discusses the wind load analysis for tank farms and other cylindrical structures, providing a comprehensive overview of the key factors, methodologies, and considerations for engineers.

Understanding Wind Loads

Wind loads are external forces acting on a structure due to the dynamic pressure and drag exerted by moving air. The magnitude and direction of wind loads depend on various factors, including:

  • Basic Wind Speed (V): This is the reference wind speed, typically determined for a specific location and return period (e.g., 50-year return period). Building codes and standards like ASCE 7-16 (“Minimum Design Loads and Associated Systems for Buildings and Other Structures”) and EN 1991-1-4 provide wind speed maps for various regions.
  • Exposure Category: This accounts for the surrounding terrain and influences the wind speed experienced by the structure. Different exposure categories are defined in codes, ranging from open terrain to urban and suburban environments.
  • Topographic Effects: Local terrain features like hills and valleys can significantly influence wind speeds and turbulence intensity.
  • Structure Shape and Size: The shape and size of the structure play a crucial role in determining the wind pressure distribution. Cylindrical structures experience wind loads differently compared to flat or rectangular structures.

Wind Load Analysis Methods for Cylindrical Structures

Several approaches can be adopted for wind load analysis of cylindrical structures:

  • Simplified Methods: Building codes often provide simplified procedures for calculating wind pressures on basic shapes like cylinders. These methods typically involve applying an equivalent static wind pressure acting on the projected area of the structure. While convenient, these methods may not be suitable for complex geometries or situations with significant topographic effects.
  • Analytical Methods: Analytical methods utilize established formulas based on wind tunnel experiments and theoretical principles to calculate wind pressures on cylindrical structures. These methods consider factors like wind speed, exposure category, and surface roughness. However, they may involve complex calculations and require specialized knowledge.
  • Computational Fluid Dynamics (CFD): This advanced method employs computational software to simulate the flow of air around the structure. CFD can generate detailed pressure distributions on the entire structure, accounting for complex geometries and local effects. However, CFD analysis requires expertise and significant computational resources.
Wind Pressure Distribution, Farm (Left) vs. Solo (Right) Tank (Source: Simscale.com)

Specific Wind Load Considerations for Tank Farms

For the calculation of wind load action effects on circular cylinder elements, the total horizontal wind force is calculated from the force coefficient corresponding to the overall effect of the wind action on the cylindrical structure or cylindrical isolated element.

The calculated effective wind pressure weff and total wind force FW correspond to the total wind action effects and they are appropriate for global verifications of the structure according to the force coefficient method. For local verifications, such as verification of the cylinder’s shell, appropriate wind pressure on local surfaces must be estimated according to the relevant external pressure coefficients, as specified in EN1991-1-4 §7.9.1.

For cylinders near a plane surface with a distance ratio zg/b < 1.5 special advice is necessary. See EN1991-1-4 §7.9.2(6) for more details. For a set of cylinders arranged in a row with normalized center-to-center distance zg/b < 30 the wind force of each cylinder in the arrangement is larger than the force of the cylinder considered as isolated. See EN1991-1-4 §7.9.3 for more details.

The calculated wind action effects are characteristic values (unfactored). Appropriate load factors should be applied to the relevant design situation. For ULS verifications the partial load factor γQ = 1.50 is applicable for variable actions.

When analyzing wind loads on tank farms, additional factors come into play:

  • Spacing and Interaction: The proximity of tanks within a farm can significantly influence wind pressures. Shielding effects and aerodynamic interaction between tanks need to be considered. Several empirical methods and CFD simulations are available to account for these effects.
  • Appurtenances: Wind loads also act on appurtenances like piping, ladders, and platforms attached to tanks. These loads can be significant and need to be included in the overall wind load analysis.
  • Dynamic Amplification: Tanks may experience dynamic amplification of wind loads due to their inherent dynamic properties. This can be particularly crucial for slender tanks or those with low natural frequencies.

Wind Load Analysis Example

A cylindrical structure of diameter (b) 5m and length (l) = 20 m is to be constructed in an area of terrain category II with a basic wind velocity vb of 40 m/s. The orientation of the cylindrical element is vertical and the maximum height above ground of the cylindrical element z = 20 m. The surface of the tank is made of galvanised steel. Calculate the wind force on the tank (Take Air density: ρ = 1.25 kg/m3)

Solution

Calculation of peak velocity pressure

The reference height for the wind action ze is equal to the maximum height above the ground of the section being considered, as specified in EN1991-1-4 §7.9.2(5). The reference area for the wind action Aref is the projected area of the cylinder, as specified in EN1991-1-4 §7.9.2(4). Therefore:

ze = z = 20 m
Aref = bl = 5 m × 20m = 100 m2

Basic wind velocity vb = 40 m/s.

For terrain category II the corresponding values are z0 = 0.050 m and zmin = 2.0 m.
The terrain factor kr depending on the roughness length z0 = 0.050 m is calculated in accordance with EN1991-1-4 equation (4.5):
kr = 0.19 ⋅ (z0 / z0,II)0.07 = 0.19 × (0.050 m / 0.050 m)0.07 = 0.19

The roughness factor cr(ze) at the reference height ze accounts for the variability of the mean wind velocity at the site.
cr(ze) = kr ⋅ ln(max{zezmin} / z0) = 0.19 × ln(max{20 m, 2 m} / 0.050 m) = 1.1384

The orography factor is considered as c0(ze) = 1.0

The mean wind velocity vm(ze)
vm(ze) = cr(ze) ⋅ c0(ze) ⋅ vb = 1.1384 × 1 × 40 m/s = 45.54 m/s

The turbulence intensity Iv(ze)
Iv(ze) = kI / [ c0(ze) ⋅ ln(max{zezmin} / z0) ] = 1.0 / [ 1.000 × ln(max{20 m, 2.0 m} / 0.050 m) ] = 0.1669

The basic velocity pressure qb 
qb = (1/2)ρvb2 = (1/2) × 1.25 kg/m3× (40.00 m/s)2 = 1000 N/m2 = 1.000 kN/m2

where ρ is the density of the air in accordance with EN1991-1-4 §4.5(1). In this calculation, the following value is considered: ρ = 1.25 kg/m3.

The peak velocity pressure qp(ze) at reference height ze
qp(ze) = (1 + 7⋅Iv(ze)) ⋅ (1/2) ⋅ ρ ⋅ vm(ze)2 = (1 + 7 × 0.1669) × (1/2) × 1.25 kg/m3 × (45.54 m/s)2 = 2810 N/m2
⇒ qp(ze) = 2.810 kN/m2

The peak wind velocity v(ze) at reference height ze 
v(ze) = [2 ⋅ qp(ze) / ρ ]0.5 = [2 × 2.810 kN/m2 / 1.25 kg/m3 ]0.5 = 67.05 m/s

Calculation of wind forces on the structure

The wind force on the structure Fw for the overall wind effect is estimated according to the force coefficient method as specified in EN1991-1-4 §5.3.

Fw = cscd ⋅ cf ⋅ qp(ze) ⋅ Aref

In the following calculations, the structural factor is considered as cscd = 1.000.

Reynolds number
Reynolds number characterizes the airflow around the object. For airflow around cylindrical objects, Reynolds number is calculated according to EN1991-1-4 §7.9.1(1):

Re = b ⋅ v(ze) / ν = (5 m × 67.05 m/s) / 15.0 × 10-6 m2/s = 22.3505 × 106
where the kinematic viscosity of the air is considered as ν = 15.0 × 10-6 m2/s in accordance with EN1991-1-4 §7.9.1(1).

Effective slenderness
The effective slenderness λ depends on the aspect ratio and the position of the structure and it is given in EN1991-1-4 §7.13(2).

For circular cylinders with length l ≤ 15 m the effective slenderness λ is equal to:
λ15 = min(l / b, 70) = min(20m / 5m, 70) = 4

For circular cylinders with length l ≥ 50 m the effective slenderness λ is equal to:
λ50 = min(0.7l / b, 70) = min(0.7 × 20 m / 5 m, 70) = 2.800

For circular cylinders with intermediate length 15 m < l < 50 m the effective slenderness λ is calculated using linear interpolation:
λ = λ15 + (λ50 – λ15) ⋅ (l – 15 m) / (50m – 15m) = 4 + (2.8 – 4) × (20 m – 15 m) / (50m – 15m) = 3.829

End effect factor
The end effect factor ψλ takes into account the reduced resistance of the structure due to the wind flow around the end (end-effect). The value of ψλ is calculated in accordance with EN1991-1-4 §7.13. For solid structures (i.e. solidity ratio φ = 1.000) the value of the end effect factor ψλ is determined from EN1991-1-4 Figure 7.36 as a function of the slenderness λ.

The estimated value for the end effect factor is ψλ = 0.658

Equivalent surface roughness
The equivalent surface roughness k depends on the surface type and it is given in EN1991-1-4 §7.9.2(2). According to EN1991-1-4 Table 7.13 for surface type “galvanized steel” the corresponding equivalent surface roughness is k = 0.2000 mm.

Force coefficient without free-end flow
For circular cylinders, the force coefficient without free-end flow cf,0 depends on the Reynolds number Re and the normalized equivalent surface roughness k/b. The force coefficient without free-end flow cf,0 is specified in EN1991-1-4 §7.9.2. The value cf,0 is determined according to EN1991-1-4 Figure 7.28 for the values of Re = 22.3505 ×106k = 0.2000 mm, b = 5.000 m, k/b = 0.000040.

The estimated value for the force coefficient without free-end flow is cf,0 = 0.803

Force coefficient
The force coefficient cf for finite cylinders is given in EN1991-1-4 §7.9.2(1) as:
cf = cf,0 ⋅ ψλ

where cf,0 is the force coefficient without free-end flow, and ψλ the end effect factor, as calculated above. Therefore:
cf = cf,0 ⋅ ψλ = 0.803 × 0.658 = 0.528

Pressure distribution for circular cylinders for different Reynolds number ranges and without end-effects

Total wind force
The total wind force on the structure Fw is estimated as:.
Fw = cscd ⋅ cf ⋅ qp(ze) ⋅ Aref = 1.0 × 0.528 × 2.810 kN/m2 × 100.00 m2 = 148.495 kN

The total wind force Fw takes into account the overall wind effect. The corresponding effective wind pressure weff on the reference wind area Aref is equal to:
weff = Fw / Aref = 148.495 kN / 100.00 m2 = 1.485 kN/m2

Note:
The effective pressure weff = 1.485 kN/m2 is appropriate for global verifications of the structure according to the force coefficient method. It is not appropriate for local verifications of structural elements, such as the shell of the cylinder. For the latter case appropriate wind pressure on local surfaces must be estimated according to the relevant external pressure coefficients, as specified in EN1991-1-4 §7.9.1.

Design Implications

The results of the wind load analysis are crucial for designing safe and efficient cylindrical structures. The wind loads are translated into equivalent static forces and moments, which are then incorporated into structural analysis software to assess the stresses and deflections in the structure. Based on these results, engineers can:

  • Determine the appropriate wall thickness and material properties for the tank shell.
  • Design roof support systems capable of withstanding wind uplift and wind-induced vibrations.
  • Optimize the anchorage system for the tank to ensure stability under wind loads.
  • Evaluate the potential need for additional bracing or wind mitigation measures.

Conclusion

Wind load analysis plays a vital role in ensuring the safety and functionality of tank farms and other cylindrical structures. Understanding the wind load characteristics, utilizing appropriate analysis methods, and considering specific complexities like tank farm interaction are crucial for engineers to design robust and wind-resistant structures. Continuous advancements in software and computational techniques are expected to further enhance the accuracy and efficiency of wind load analysis in the future.

error: